Sunteți pe pagina 1din 15

Biochimica et Biophysica Acta 1861 (2017) 824–838

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

journal homepage: www.elsevier.com/locate/bbagen

Review

Minor snake venom proteins: Structure, function and


potential applications
Johara Boldrini-França a, Camila Takeno Cologna a, Manuela Berto Pucca b, Karla de Castro Figueiredo Bordon a,
Fernanda Gobbi Amorim a, Fernando Antonio Pino Anjolette a, Francielle Almeida Cordeiro a,
Gisele Adriano Wiezel a, Felipe Augusto Cerni a, Ernesto Lopes Pinheiro-Junior a, Priscila Yumi Tanaka Shibao a,
Isabela Gobbo Ferreira a, Isadora Sousa de Oliveira a, Iara Aimê Cardoso a, Eliane Candiani Arantes a,⁎
a
School of Pharmaceutical Sciences of Ribeirão Preto, University of São Paulo, Ribeirão Preto, Brazil
b
Medical School of Roraima, Federal University of Roraima, Boa Vista, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Snake venoms present a great diversity of pharmacologically active compounds that may be applied as research
Received 20 August 2016 and biotechnological tools, as well as in drug development and diagnostic tests for certain diseases. The most
Received in revised form 12 December 2016 abundant toxins have been extensively studied in the last decades and some of them have already been used
Accepted 20 December 2016
for different purposes. Nevertheless, most of the minor snake venom protein classes remain poorly explored,
Available online 22 December 2016
even presenting potential application in diverse areas. The main difficulty in studying these proteins lies on
Keywords:
the impossibility of obtaining sufficient amounts of them for a comprehensive investigation. The advent of
Snake venoms more sensitive techniques in the last few years allowed the discovery of new venom components and the in-
Toxins depth study of some already known minor proteins. This review summarizes information regarding some struc-
Drug discovery tural and functional aspects of low abundant snake venom proteins classes, such as growth factors, hyaluroni-
dases, cysteine-rich secretory proteins, nucleases and nucleotidases, cobra venom factors, vespryns, protease
inhibitors, antimicrobial peptides, among others. Some potential applications of these molecules are discussed
herein in order to encourage researchers to explore the full venom repertoire and to discover new molecules
or applications for the already known venom components.
© 2016 Published by Elsevier B.V.

1. Introduction compose the venom directly reflects its toxicity and pathophysiologic
effects, and may represent an evolutionary arms race, in which the
Snake venoms are complex mixtures of organic and inorganic com- venom mixture is adapted to improve the predator's ability of subduing
pounds that act on a variety of specific metabolic and physiologic targets different preys and to overcome the resistance of some prey species to
of preys or victims, assisting feeding and defense [1]. Protein composi- the venom [5,6].
tion of snake venoms may undergo pronounced qualitative and quanti- Elapidae and Viperidae venom proteins are produced and secreted
tative variation in all levels of taxa, as well as in populations and by oral exocrine glands that present a basal-central lumen, where the
individuals of the same species [2,3]. Variation in protein expression of produced venom is stored until the moment of its delivery [7]. The pro-
venom components may also be observed in the same specimen in dif- duction of toxins is activated by morphological and biochemical chang-
ferent ontogenetic stages [4]. Such great diversification in proteins that es in secretory epithelial cells after venom injection or extraction [8–
10]. However, synthesis and secretion of different protein classes are
not synchronized and may result in variation in venom composition in
⁎ Corresponding author at: Universidade de São Paulo, Faculdade de Ciências different stages of the venom production cycle [11].
Farmacêuticas de Ribeirão Preto, Departamento de Física e Química, Av. do Café s/n°,
Many regulatory mechanisms are supposed to impact protein com-
Monte Alegre, 14040–903 Ribeirão Preto, SP, Brazil.
E-mail addresses: joharafran@gmail.com (J. Boldrini-França), camilatcbio@gmail.com position in snake venoms, such as mutations affecting gene expression
(C.T. Cologna), manupucca@hotmail.com (M.B. Pucca), karla@fcfrp.usp.br [12], duplication and loss of toxin-related genes [13], post-transcrip-
(K.C.F. Bordon), fernandagamorim@gmail.com (F.G. Amorim), fanjolette@yahoo.com.br tional microRNAs regulation [14] and proteolytic processing [13]. Al-
(F.A.P. Anjolette), fran_acordeiro@hotmail.com (F.A. Cordeiro), gisele.wiezel@gmail.com though presenting variable expression levels in different snake
(G.A. Wiezel), felipe_cerni@hotmail.com (F.A. Cerni), ernesto.pinheiro@usp.br
(E.L. Pinheiro-Junior), priscila.shibao@gmail.com (P.Y.T. Shibao),
venoms, the pathologically important toxins, such as phospholipases
igobboferreira@yahoo.com.br (I.G. Ferreira), isadora_so@yahoo.com (I.S. de Oliveira), A2, metalloproteases, kallikrein toxin families and three-finger toxins
iara_cardoso@hotmail.com (I.A. Cardoso), ecabraga@fcfrp.usp.br (E.C. Arantes). are usually the most abundant proteins found in these biological

http://dx.doi.org/10.1016/j.bbagen.2016.12.022
0304-4165/© 2016 Published by Elsevier B.V.
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 825

samples. Some of these toxins compose up to 80% of the total venom evolved for a more comprehensive meaning and nowadays is also used
proteins, such as phospholipase A2 from Crotalus durissus venom, to describe the studies which combine strategies such as genomics,
which forms the crotoxin complex that is a neurotoxin responsible for transcriptomics and proteomics to unveil the full picture of venom com-
the major clinical symptoms of crotalic envenoming in South America ponents [25,28]. A genomic analysis of the genes encoding venom pro-
[15]. The major snake toxins are encoded by large multilocus gene fam- teins may answer a number of fundamental questions, especially
ilies that exhibit substantial gene duplication and directional selection regarding the evolutionary origin of those remarkable compounds and
and consequently the greatest variation [13]. the genetic basis of generating toxin function and diversity [28,29]. In
On the other hand, the minor snake venom proteins are supposed to defiance of this great value, there is still a lack of snakes' full genomes
have ancillary functions, less variability and experienced little to no in the literature [29,30]. The big size of snakes' genomes which could
gene duplication and positive selection. This contrast may be a result turn the assembling of novel genomes an extreme difficult task could
of the likely conservative functions and the low relative abundance of explain this deficiency [28]. In a recent study regarding the loss of
ancillary proteins in the venoms, which precluded their participation venom toxins genes in rattlesnakes, Dowell and colleagues [29] stated
in the pray-predator arm's race, resulting in a lack of selective pressures that most investigations of snake venom diversity have been conducted
[13,16]. Nevertheless, although some minor snake venom protein clas- without genome sequences; besides their article, only other two groups
ses are supposed to present secondary effects, their ecological relevance reported full genomes from snakes [30].
to the snake and their physiologic effects should be better explored. It is Following a different direction, transcriptomics, essentially the next
also worth to mention that most of these proteins interact with specific generation sequencing (NGS) transcriptomic approach, have allowed
targets in some physiological systems, which will be discussed thereaf- the investigation of the venom repertoire in a broader way, revealing
ter. Accordingly, these features have aroused an interest in investigating transcripts that are rarely expressed or present in very low abundance
possible medical and biotechnological applications for these families of and therefore poorly studied hitherto [24,31]. Nevertheless, such strat-
venom proteins. egy cannot always provide accurate quantitative data nor predict the
Another issue to be considered is that venom production has a sub- post translational modifications (PTM), which include the cleavage of
stantial metabolic cost [17] and a long regeneration cycle [9,18], which the mature region, disulfide bonds formation, modifications of the side
indicates that the venom composition is optimized in order to maximize chains and/or N- and C-terminal [25,32,33]. Glycosylation, as an exam-
efficacy while metabolic expenditure is minimized [19,20]. Thus, the ple, is one of the most frequent modifications found in snake venoms
synthesis and secretion of this variety of protein classes presenting and play notable role in protein folding, conformation, stability, phar-
merely ancillary and unessential functions seem contradictory consid- macodynamics, and biological activity. As for other types of PTMs, glyco-
ering that the maintenance of the venom repertoire is considerably sylation increase the complexity of venom proteomes and expand
metabolic expensive. Therefore, researchers should look into minor pro- functions of its components [34], highlighting the relevance to elucidate
teins classes in the venom aiming to unveil possible functions they may those protein adornments [33,35]. In parallel and to fulfill those gaps,
play in the envenoming or in the physiology of the venom gland. More- proteomics, supported by advances of mass spectrometry, has been
over, the expression of these proteins in nonvenomous tissues, the con- widely employed to explore animal venoms. This strategy can elucidate
servation degree of physiologic activities comparing with their related not only the molecular mass of the toxin, but also fragments of its se-
orthologs and the structure and rate of duplication, mutations and rear- quence which can be used as tags for database searches, the identifica-
rangement of their encoding genes may provide relevant information tion of post translational modifications and their localization within the
regarding snake venom evolution and the generation of toxins diversity. sequence [36].
In spite of their possible great scientific value, some of these classes Individually, proteomics and transcriptomics strategies have already
are poorly explored mainly due to difficulties in purifying them from the contributed massively to the investigation of snake venom diversity
venom for an in-depth investigation [21]. In this context, this review (Table 1). Notwithstanding, the “venomics” strategy, which is the inte-
will focus on some classes of proteins that are known to be expressed gration of the data coming from different cutting-edge technologies,
in relatively low amounts in most snake venoms (b 20% of the whole will replace the classical approach, giving toxinologists the access be-
venom proteins, as showed in Table 1), such as growth factors, hyal- yond the traditional and paving the way to explore bona fide novel com-
uronidases, cysteine-rich secretory proteins, nucleases and nucleotid- pounds and minor venom proteins [37–39].
ases, cobra venom factors, vespryns, protease inhibitors, antimicrobial
peptides, among others. Moreover, this review also aims to discuss the 3. Minor venom protein classes
importance of minor proteins to toxinology and how the advent of inno-
vative technological resources may enable the assessment of their struc- 3.1. Growth factors
tural and functional properties and their role in the evolutionary history
of snake venoms. We also highlight that this review represents the first The designation “growth factor” was historically associated with
compilation of snake venom minor proteins and provides useful infor- growth and cell proliferation. Later, other cellular responses were attrib-
mation of those still neglected molecule that may hide outstanding po- uted to these neurotrophins, such as cell differentiation, transformation,
tential applications. synthesis, secretions, death and motility [105]. The first identified
growth factor was a nerve growth factor (NGF) from mouse sarcoma
2. Omic approaches in the identification of minor venom proteins 180 in the 1950s [106,107]. Snake venom NGF (sv-NGF) was firstly iso-
lated in 1956 from the venom of Agkistrodon piscivorus [108]. Some
Until a few years ago, comprehensive investigation of snake venoms years later, epidermal growth factor (EGF) from the submaxillary
composition was considered a time-costing and laborious work. Mainly gland of mouse and platelet-derived growth factor (PDGF) from mon-
due to the low sensitivity of the applied methods, usually only the major key blood were discovered [109,110]. Since then, over 200 growth fac-
components, such as proteases, neurotoxins and phospholipases A2, tors have been isolated and studied [105].
were studied in detail [22,23]. The advances of proteomics, Among the several known families of growth factors, the following
transcriptomic and genomic techniques, as well as the improvement have been identified in snakes and are deposited in UniProt (Universal
of their related instrumentation, have led to progress in knowledge re- Protein Resource Knowledgebase) databank: latent-transforming
garding the complexity of snake venom composition and the evolution- growth factor beta-binding protein (LTBP), transforming growth factor
ary history of those sophisticated lethal cocktails [24,25]. Not long ago, beta-2 (TGF-β), hepatocyte growth factor-regulated tyrosine kinase
the term “venomics” was used for the first time to describe the proteo- (HGF), hepatoma-derived growth factor (HDGF), delta/notch-like epi-
mic study of snake venom composition [26,27]. However, the term has dermal growth factor, connective tissue growth factor, insulin-like
826 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

Table 1
Main minor protein classes in snake venoms.

Venom protein classes Approaches Relative Main biological activities Other relevant characteristics References
used in the abundance in
identification snake venoms (%)⁎

5’-Nucleotidases T, P and T.A. b0,1 - 4,8 Hydrolysis of 5'-nucleotides to Act in synergism with ADPases, [40–42]
nucleosides, leading to platelet phospholipases, and disintegrins,
aggregation inhibition potentiating the anticoagulant effect in
the envenoming
Aminopeptidase A T, P and T.A. - Hypotension - [43–45]
Aminopeptidase N P - Hypotension - [44]
Cathelicidin-derived and
T -- Antimicrobial activity -- [46–57]
cathelicidin-related peptides
Cobra venom factors (CVFs) P, T And T.A. 0.1 – 2.8 Complement depletion Mainly found in Elapidae snakes [58–62]
Cystatins T, P and T.A. 1,7 Protease inhibition Inhibition of tumor cells metastasis [63–65]
Presence of sixteen cysteine residues
Cysteine-rich secretory proteins
P, T and TA 0,1 - 15,9 Blockage of ion channels highly conserved which form eight [66–70]
(CRISPs)
disulfide bonds
Dipeptidylpeptidase IV P and T - Alterations in the cardiovascular, - [44,45,71,72]
immune and neuroendocrine systems
and glucose hemostasis
Act as a “spreading factor” in the
Degradation of hyaluronan in
Hyaluronidases P, T and TA 0.1 – 1.9 envenoming and potentiate the effects of [73–75]
extracellular matrix
toxins
Kazal-type inhibitors P, T 8,3 Protease inhibition -- [76,77]
Diversity of biological effects, including
P, T and blockage of ion channels, disturbances in
kunitz-type inhibitors 0,3 - 12,6⁎⁎ Protease inhibition [78–87]
TA blood coagulation, fibrinolysis and
inflammation
Increases the vascular permeability and
Neuronal differentiation, synaptic
facilitates the diffusion of toxins in
Nerve growth factors (NGFs) P, T and TA 0.1 – 5 plasticity, and neuroprotection in the [88–90]
envenoming. May act as a proapoptotic
peripheral and central nervous systems.
factor
Hypotension, locomotor depression and
Phosphodiesterases (PDEs) P, T and T.A. 0.1 – 3,2 -- [91–94]
platelet aggregation
Cleavage of membrane phospholipids Hydrolysis of phospholipids at sn-1 and
Phospholipases B (PLB) P, T and TA b0.1 – 1,4 [95–97]
and hemolytic activity sn-2 positions
Vascular endothelial Proliferation of vascular endothelial cells
P, T and TA b0.1 – 3,2 -- [98–100]
growth factors (VEGFs) and hypotension
Vespryns T, P and TA 0.2 –14,4 Hypolocomotion, hypernociception Ohanin presents a single cysteine residue [101]
Structural similarity to whey acidic
Waprins T and TA -- Antimicrobial activity [102–104]
protein (WAP)

T = transcriptome; P = proteome; T.A. = traditional approach (activity-guided assays).


⁎ Relative abundance determined by proteomic approach.
⁎⁎ Exception of Dendroaspis polylepis venom that is composed by 61.1% of kunitz-type inhibitor.

growth factor-binding protein 5 (IGF), epidermal growth factor (EGF- F, the sv-NGFs from Acanthophis antarcticus, Anilios nigrescens (formerly
like protein), brain-derived neurotrophic factor (BDNF), interferon Ramphotyphlops nigrescens) and Oxyuranus scutellatus were confirmed
(IF), platelet-derived growth factor (PDGF), tumor necrosis factor to be N-glycosylated proteins [114]. Notably, this glycosylation site is
(TNF), snake venom vascular endothelial growth factor (sv-VEGF) and missing in Notechis scutatus, Pseudechis australis and Pseudechis
nerve growth factor (NGF). However, this review will focus on VEGF porphyriacus [114]. Although the sv-NGFs from the snakes Hoplocephalus
and NGF, which are the most studied snake venom growth factors. stephensii [114] and Naja sputatrix [115] have shown an N-linked glyco-
sylation site, it migrates as a non-glycosylated protein. The functional rel-
3.1.1. Nerve growth factors (NGFs) evance of these differences in PTMs for sv-NGF demands further assays.
NGFs participate in neuronal differentiation, synaptic plasticity, and Sv-NGF has low abundance in snake venoms, corresponding to
neuroprotection in the peripheral and central nervous systems [88]. 0.1%–0.5% (w/w) of crude venoms of Daboia russelii (formerly Vipera
NGF acts also on no neuronal cells, especially on hematopoietic stem russelli) [116] and of snakes from Oxyuranus [112,114], Naja [90,117,
cells. However, most of these effects were reported for murine NGF 118] and Sistrurus [119] genus. The relative abundance of ESTs coding
[111]. sv-NGF comprised only 0.1% of the total transcripts from Lachesis muta
The physiological responses to sv-NGF have not been elucidated yet venom gland [98].
[111]. However, the effects of mammalian NGF on non-neuronal cells The cDNA encoding sv-NGF exhibits a signal peptide, a pre-prodomain
might assist the comprehension of some effects exerted by sv-NGFs on and the mature protein [115]. The precursor contains a presumptive 18
mammalian systems [112]. It has been described that part of the sv- amino acid signal sequence, which is followed by a propeptide (proNGF)
NGF injected at the bite site may reach the circulation, leading to of approximately 109 amino acid residues [120]. The prodomain is impli-
some physiological activities on non-neuronal cells or tissues. Sv-NGF cated to the correct folding of mature NGF [112]. The first cDNA of a sv-
from Naja atra (formerly N. naja atra) exerts important systemic effects NGF was reported in 2002 [121] from Bothrops jararacussu venomous
in the envenoming, such as plasma extravasation and histamine release, gland. The molecular model revealed that this mature sv-NGF (containing
which result in tissue vulnerability and facilitate toxin diffusion in the 118 amino acid residues) is formed by a pair of β-sheets, three β-hairpin
prey organism [111]. In contrast, the sv-NGF isolated from Naja naja loops, a reverse turn and a short α-helix.
has shown no toxic effects [113]. Concerning the functional properties and potential applications of
In regards to the presence of PTMs, some sv-NGFs were confirmed to sv-NGFs, the injection of cobra venom NGF (cvNGF) was capable of
have an Asn23 glycosylation site (NXS/T). After treatment with PNGase slow down the growth of Ehrlich ascites carcinoma (EAC) cells in
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 827

mice, probably via an indirect mechanism in which tyrosine kinase A β → 1–4 of the residues N-acetyl-β-D-glucosamine and D-glucuronate
(TrkA) receptors are involved. On the other hand, cvNGF showed prolif- from hyaluronan producing tetra and hexasaccharides [74]. Therefore,
erative activity and had no cytotoxic effect on breast cancer cell line during the envenoming, hyaluronidases facilitate the venom diffusion
MCF-7 [122]. in the victim's tissue due to their hydrolytic characteristics, acting as a
Sv-NGF from N. sputatrix upregulates endogenous expression of NGF “spreading factor” and enhancing the toxins' effects [75,134,135]. This
in PC12 cells, pro-survival cell surface receptors and ion channels [115]. class of enzyme has been found in several organisms, being ubiquitous
The neurite differentiation activity evidenced with sv-NGF from N. in snake venoms [73,136,137].
sputatrix [115] and O. scutellatus [114] was similar to that of mammalian Hyaluronidases from different sources have been used in different
homolog. These results may lead to further studies on sv-NGFs as poten- applications in medicine. Examples include the use as a diffusion pro-
tial therapeutic agents against neuronal injury. moter for active substances (drugs, for instance), the treatment of
hyaluronan-induced diseases (such as some types of cancers), and in
3.1.2. Vascular endothelial growth factor (VEGF) the aesthetic medicine [138]. Despite their great therapeutic potential,
Vascular endothelial growth factor (VEGF), formerly designated as these enzymes are found in small proportions in snake venoms and
vascular permeability factor, stimulates vasculogenesis, angiogenesis they have an extremely unstable catalytic activity. These issues hamper
or lymphangiogenesis. The VEGF family is divided into seven groups, their isolation from venoms, as well as their in-depth functional and
designated as VEGF-A to VEGF-F and placental growth factor (PGF) structural characterization [139,140]. This is reinforced by the few hyal-
[99]. Snake venom glands present at least three different VEGF-Fs with uronidase amino acids sequences deposited in databases, which are cur-
unique features and with distinct receptor-selectivity, designated as rently fourteen for Uniprot and seventeen for NCBI.
VEGF-F1 to VEGF-F3. Additionally, VEGF-A-like transcripts have also The first report indicating the occurrence of hyaluronidase in snake
been identified in some snake venoms [123]. VEGF-A is a cytokine se- venoms was in 1939 by Duran-Reynalds, who analyzed the hyaluroni-
creted by tumor cells that play important role in normal and tumor-re- dase activity in the venom of nine different snake species [134]. Howev-
lated angiogenesis. er, up to date, limited studies have successfully isolated these enzymes
Up to now, 26 expressed sequence tags (ESTs) for sv-VEGF were from snake venoms. Reports include the hyaluronidases from
identified by transcriptomic approach in the venom glands of Deinagkistrodon acutus (formerly Agkistrodon acutus) [141], Agkistrodon
Bothropoides insularis (formerly Bothrops insularis) (19), Bothropoides contortrix [142], N. naja [136], D. russelli [143], Cerastes cerastes [135],
pauloensis (former Bothrops pauloensis) (1), Bothrops atrox (former Crotalus durissus terrificus [137] and Lachesis muta rhombeata [140].
Bothrops colombiensis) (3), Crotalus durissus collilineatus (2) and L. Snake venom hyaluronidases are glycoproteins with molecular mass
muta (1) snakes [98,124–126]. The sv-VEGFs corresponded to 0.1% of ranging from 28 to 70 kDa, with optimum enzyme activity detected
the total ESTs from L. muta venom gland [98]. The VEGF gene is com- around pH 5.5, 37 °C and 0.15–0.2 M NaCl. They hold specificity for
prised of eight exons which yields the splicing isoforms VEGF-An, hyaluronan and show no significant activity upon chondroitin sulfate
where “n” may be 121, 145, 165, 183, 189 or 206 and represents the [135,136,140–143]. Moreover, hyaluronidase activity can vary accord-
number of amino acid residues in the protein [99]. VEGF-A165 is the ing to the snake's age [144], species [145,146] and habitat [147,148].
major isoform and a biological indicator of the invasiveness of hepato- The enzyme activity is inhibited by divalent cations, high temperatures
cellular carcinoma [127]. In 1999, a VEGF-like protein, named HF, was and salt concentrations, extreme pH and by EDTA, heparin, denaturing
firstly isolated from Vipera aspis aspis venom [123,128]. A cDNA agents and herbal extracts [135,137,142,143].
encoding for sv-VEGF was identified in B. insularis venom gland in As previously mentioned, in vivo experiments have proven that
2001 [129]. Two sv-VEGFs (vammin and VR-1 from Vipera ammodytes snake venom hyaluronidase acts as a spreading factor, enhancing the ef-
ammodytes and Daboia russelli russelli venoms, respectively) are fect of venom toxins [135–137,143]. Therefore, many recent studies put
known to lack the C-terminal heparin-binding region found in other a spotlight on the production of antibodies and on the search for syn-
heparin-binding VEGF subtypes. On the other hand, their C-terminal thetic and natural inhibitors of these enzymes in order to improve anti-
recognizes similar heparin/heparan sulfate molecules and shows high venom therapy, since the importance of hyaluronidases in the
selectivity for the kinase insert domain-containing receptor (KDR) envenoming process is outstanding [135,136,149–155].
when compared with VEGF-A165. Besides that, their C-terminal specif-
ically blocks the VEGF-A165 activity [99,130].
VEGF-F induces hypotension and proliferation of vascular endotheli- 3.3. Cysteine-rich secretory proteins
al cells by binding to the KDR receptor [100]. VEGF-A165 and vammin
induced the synthesis and secretion of perlecan via VEGF receptor-2 Cysteine-rich secretory proteins (CRISPs) are non-enzymatic com-
(VEGFR-2) in cultured human brain microvascular endothelial cells. ponents present in various organisms [66]. They are also found in
That proteoglycan may help to regulate the angiogenesis, maintain the snake venoms, but their function in envenoming has not been fully un-
endothelial barrier function and contribute to a synergistic effect on an- derstood thus far [156]. Snake venom CRISPs are single chain proteins
giogenesis [131]. The activation of VEGFR-2 depends on the affinity and with molecular masses ranging from 20 to 30 kDa. They display sixteen
concentration of the ligands and the angiogenic activity is potentiated highly conserved cysteine residues that form eight disulfide bonds [66].
when the ligand binds to the co-receptor Neuropilin (Nrp). Hence, CRISPs are distributed among Viperidae and Elapidae families from
drugs with the VEGFR-2/Nrp binding domains may be successfully de- different continents. There are reports of the isolation and cloning of
signed as potential proangiogenic therapies [132,133]. However, further three snake venom CRISPs from Agkistrodon piscivorus piscivous,
studies are necessary to elucidate the biological activities, mechanism of Ophiophagus hannah and Crotalus atrox, named piscivorin, ophanin
action and possible applications of these neurotrophins from snake and catrin, respectively [157]. Other CRISPs such as triflin, ablomin,
venoms. latisemin and tigrin were isolated from the venoms of Protobothrops
flavoviridis, Agkistrodon blomhoffi, Laticauda semifasciata and Rhabdophis
3.2. Hyaluronidases tigrinus tigrinus, respectively. Ablomin, triflin and latisemin block the
caffeine-induced depolarization, but they have no effects on the con-
Venom hyaluronidases are enzymes that hydrolyze preferentially traction of the smooth muscle, indicating their L-type Ca2 + channel
the hyaluronan, the major glycosaminoglycan found in the extracellular blocker action. On the other hand, tigrin does not inhibit depolarization
matrix [73]. These enzymes belong to EC 3.2.1.35 class, which includes and contraction induced by caffeine [67]. Pseudechetoxin (PsTx) and
the hyaluronidases found in mammals' sperm cells, lysosomes and ani- pseudecin, from P. australis and P. porphyriacus venoms, respectively,
mals venoms. They are capable of hydrolyzing glycosidic linkages block the cyclic nucleotide-gated (CNG) ion channel [67,68].
828 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

Natrin, from N. a. atra venom, was reported to be a BKCa channel hydrolyze DNA and RNA, respectively. Exonucleases comprise phospho-
blocker [69]. Its three-dimensional structure, determined by X-ray crys- diesterases (PDEs) [94] that catalyze the hydrolysis of phosphodiester
tallography, revealed three functional sites: a pathogenesis-related pro- bonds, releasing 5′-mononucleotides [91].
tein from the group 1 (PR-1) like domain localized at the N-terminal Endonucleases show optimum activity on acidic pH and do not re-
region, a highly flexible cysteine-rich domain (CRD) at the C-terminal quire divalent cations for nucleic acids hydrolysis. On the other hand,
extremity of the protein, and a hinge between these domains. The PDEs have optimum activity on basic pH and divalent metal ions are re-
CRD domain plays an important role in blocking the ion channel activity quired for PDE's hydrolytic activity [167–169].
since it probably interacts with the pore of Kv1.3 channels by hydrogen Endonucleases and PDEs have been found in snake venoms from
bonds formation [70]. Elapidae family and crotaline and viperine snakes from Viperidae family
Besides blocking the ion channel activity, CRISPs may act as inflamma- [71]. The biological activities of endonucleases have not been elucidated
tory modulators [158] once they can regulate the expression of adhesion yet [91] Nevertheless, PDEs are known to lead to hypotension, locomo-
molecules in endothelial cells. Lecht et al. reported that a CRISP from Echis tor depression [92] and inhibition of platelet aggregation [93].
carinatus sochureki negatively regulates angiogenesis, thus disturbing the In general, nucleases can lead to the production of purine and
wound healing process of envenomed victims [159]. Furthermore, pyrimidine nucleotides that might act synergistically with nucleotidases
crovirin, a CRISP purified from Crotalus viridis viridis [160] was tested in and others snake venom toxins (phospholipases A2, cardiotoxins,
vitro for its antiparasitic activity. This CRISP demonstrated a dose-depen- myotoxins, cytolytic peptides and proteases/hemorrhagins) [71,91,
dent antiparasitic activity on Leishmania amazonensis and Trypanosoma 161,170–173], contributing to the symptomatology and overall pathol-
cruzi amastigote forms and Trypanosoma brucei rhodesiense metacyclic ogy caused by the snake bite envenoming [37].
and blood trypomastigotes form.
In the light of the above, CRISPs are a promising class of snake venom 3.4.2. Nucleotidases
component that can be employed as a model for future cancer drugs due Nucleotidases are enzymes involved in the cleavage of nucleic acid
to the negative regulation of angiogenesis. In addition, this class of com- derivatives and nucleic acid-related substrates (ATP, ADP and AMP).
ponents may assist the development of new drugs against leishmaniosis However, specific identification of snake venom nucleotidases is im-
and Chagas disease. paired by the presence of other toxins that share similar substrate spec-
ificity [174].
3.4. Nucleases and nucleotidases Nucleotidases have been found in a great number of snake venoms
[168], but there are only few nucleotidases isolated and characterized
Nucleases and nucleotidases are hydrolytic enzymes ubiquitously dis- so far due to their minor amounts in venoms [174,175]. The biological
tributed in snake venoms. Their main function is related to the generation activities of nucleotidases are known to be involved in the digestive pro-
of adenosine through their catalytic activity, as shown in Fig. 1, which cess and in endogenous purines release, potentiating the venom-in-
schematically illustrates adenosine production by these enzymes. Adeno- duced hypotension and paralysis via purine receptors. However,
sine may support toxin biodistribution and also contributes to prey nucleotidases activities need to be deeper investigated [71,161]. Fur-
immobilization, whereas it increases vascular permeability, inhibits neu- thermore, some reports in the literature show that nucleotidases can
rotransmitter release and promotes hypotension, sedation, locomotor de- act independently or in synergism with other toxins, like phosphodies-
pression and bradycardia. Although nucleases and nucleotidases activities terase, phosphomonoesterase, endonuclease, phospholipases A2, cyto-
are broadly detected in snake venoms, the biological effects of these en- toxins, myotoxins, and heparinases [42,71,161,176–178].
zymes are not completely clear [45,71,161–163]. Snake venom nucleotidases comprise 5′-nucleotidase, ATPase and
ADPase. 5′-Nucleotidase are metalloenzymes of high molecular mass
3.4.1. Nucleases [42,168,179,180], ranging from 73 to 100 kDa [42,177,181,182]. Gulland
Nucleases are hydrolytic enzymes that cleave nucleic acids and their and Jackson [183] firstly identified the 5′-nucleotidase activity in snake
derivatives [91]. They are classified as endonucleases and exonucleases. venoms. These enzymes selectively hydrolyze 5′-nucleotides to nucleo-
Endonucleases [164] comprise DNAses [165], and RNAses [166] which sides [180] and inhibit platelet aggregation by adenosine release [40–

Fig. 1. Schematic representation of adenosine production/generation by snake venom toxins. During the envenoming, toxins (in green) promote cellular lysis, releasing nucleic acids (DNA
and RNA). Nucleic acids become susceptible to the action of snake venom nucleases, resulting in the production of 5′-nucleotides, which are cleaved by snake venom phosphatases,
generating adenosine. Another pathway to produce adenosine involves the cleavage of ATP by ATPase or phosphodiesterase, producing ADP and AMP, which are degraded by
nucleotidases, yielding adenosine.
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 829

42]. Thus, 5′-nucleotidase increases the anticoagulant effect of ADPases, of inhibiting enzymes involved in the blood coagulation [84–87]. Thus,
phospholipases A2 and disintegrins by acting synergistically with these due to their extensive biological roles, these venom proteins may pres-
toxins [184]. ent promising biotechnological and pharmaceutical applications.
ATPases were firstly described by Zeller in 1950 and are known to Other minor classes of protease inhibitors found in snake venoms
hydrolyze ATP, releasing adenosine and pyrophosphate. Depending on are the cysteine protease inhibitors (cystatins), the metalloproteases in-
the reaction conditions, ATPases can cleave ATP into AMP and pyro- hibitors and the kazal-type protease inhibitors. The latter was only re-
phosphate or into ADP and phosphate. ATPases have not been isolated ported in Bothriechis ssp. venom, with a relative abundance of 8.3 and
from snake venoms yet, and their biological activity remains unclear. 9% in B. schlegelii and B. supraciliaris species, respectively [76,77]. Re-
However, they may be involved in shock symptoms by depletion of cently, a kazal-type inhibitor-like protein that is able to inhibit trypsin
ATP [185], since ATP released from skeletal muscles cells (due to activity activity at pH 5.4 was purified from B. schlegelii venom. This finding sug-
of myotoxins) is quickly hydrolyzed by ATPase, and act as a danger sig- gests that such protein may act controlling venom proteolysis, consider-
nal molecule stimulating purinergic receptors, contributing to the path- ing that the venom gland has an acidic pH [196].
ophysiology of envenomation [186–188]. Metalloprotease inhibitors were found by proteomic approaches in
ADPase hydrolyzes ADP, producing adenosine and orthophosphate snake venoms from Echis ocellatus, C. c. cerastes [209] and V. anatolica
[168,189]. The only ADPase from snake venom isolated so far has [206]. Recently, a new group of metalloprotease peptide inhibitors
94 kDa and it was purified from D. acutus. This toxin inhibits platelet ag- (Poly-Gly- and poly-His-poly-Gly-peptides and tripeptides) was detect-
gregation in platelet rich plasma induced by different molecules (ADP, ed in snake venoms. Tripeptides were identified in the venom of
collagen and sodium arachidonate), but it does not inhibit thrombin-in- Bothrops ayerbei [210], Bothriechis supraciliaris [77] and Bothrocophias
duced aggregation in platelet poor plasma [177]. The inhibition of plate- campbelli [211]. Interestingly, these inhibitors are encoded by the
let aggregation is also attributed to adenosine release due to ADPase same precursor of bradykinin potentiating and C-natriuretic peptides.
activity [174]. Furthermore, metalloprotease inhibitors such as Poly-His-poly-Gly-
peptides (pHpG peptides) were identified in the venoms of Bothrops
3.5. Protease inhibitors snakes [212,213] and E. ocellatus [209], whereas poly-Gly-peptides (pG
peptides) were found in the venom of Bothriechis supraciliaris [77].
Snake venoms are interesting sources of protease inhibitors, al- Regarding snake venom cystatins, the first toxin of this class was iso-
though these molecules represent, in most cases, only a small propor- lated from Vipera ammodytes by Kregar and colleagues, in 1981 [214].
tion of snake venoms. The control of most body's chemical reactions is The only proteomic study that determined the percentage of this class
sustained by the antagonism of: (i) proteases, which play key functions of toxin in snake venom revealed that it represents 1.8% and 9.8% of
in different systems and biochemical pathways, and (ii) the correspon- total venom proteins of the species Bitis gabonica gabonica and Bitis
dent protease inhibitors, responsible for controlling protease activity arietans, respectively [198]. Moreover, three other cystatins were puri-
[190]. In this context, the possible function of protease inhibitors in fied from snake venom, all of them belonging to type-2 cystatins,
snake venoms is to promote an imbalance in prey homeostasis, since which present about 120 amino acid residues and two characteristic di-
these proteins can inhibit some classes of enzymes, like serine proteases sulfide loops [215–217]. Nonetheless, no studies have elucidated their
[191,192], acetylcholinesterase (AchE) [193] as well as interact with ion role in the envenoming process yet.
channels [194,195], thus interfering in the physiological systems medi- Recently, the pharmaceutical application of a recombinant snake
ated by these components. On the other hand, it was suggested that venom cystatin (sv-cystatin) in cancer therapy was investigated. The re-
these protease inhibitors may also assist the proteolysis control while combinant cystatin is capable of inhibiting the invasion and metastasis
the venom is stored at the venom gland [79,196]. of tumor cells, both in in vitro and in vivo assays [64]. Furthermore, it
The first protease inhibitor found in snake venom was reported in was reported that this molecule is capable of inhibiting tumor angiogen-
1972 when Takahashi and colleagues [197] isolated a potent kunitz- esis and tumor-endothelial cell adhesion [63,65]. Additionally, a
type protease inhibitor from the venom of D. russelli. The occurrence recombinant adenovirus transfected with the sv-cystatin gene has dem-
of this family of protease inhibitors is reported in Elapidae and Viperidae onstrated the ability to inhibit the invasion and metastasis of several
snake venoms [79,198–200]. types of tumors, such as mouse melanoma cells, human gastric carcino-
This class of protease inhibitors presents a conserved motif found in ma cells, and MHCC97H cells. This recombinant adenovirus approach
kunitz bovine pancreatic trypsin inhibitor, which is the active domain demonstrated an increased antitumor activity compared to the recom-
responsible for the inhibition of proteases such as trypsin, chymotryp- binant protein itself [63].
sin, elastase, thrombin, and activated factor X [201,202]. Structurally, Protease inhibitors are becoming molecules of interest to biotechno-
they display approximately sixty amino acid residues, with six con- logical and pharmaceutical areas since proteases are one of the main
served cysteine residues involved in three disulfide bonds, supporting potential drug targets. There are about 553 human gene products incor-
the stable and compact structure of the folded peptide [203–205]. porating protease domains. In addition, these enzymes are also found in
These venom components represent from 0.3% to 4.6% of total pro- bacteria, virus and parasites [218,219]. In this context, the search for
teins in snake venoms [198,206–208]. The exceptions identified thus new protease inhibitors plays a key role to warrant the discovery of
far are the species Dendroaspis polylepis and D. angusticeps, whose more effective drugs for human therapeutics, especially for cancer
venom proteome presented larger amounts of kunitz type protease in- treatment.
hibitors [163]. Although these inhibitors may exhibit several structural
similarities, they show an extensive range of biological functions, in-
cluding blockage of ion channels, disturbances in blood coagulation, in- 3.6. Antimicrobial peptides
flammation and fibrinolysis [78–81].
A kunitz-type protease inhibitor isolated from Macrovipera lebetina Antimicrobial peptides are found in several organisms and are part of
venom displayed a remarkable anti-tumor activity by the interaction their innate-immune system, participating of host defense and contrib-
with an integrin receptor [82]. In a further study, this molecule inhibited uting to the survival of these organisms in microbe-rich environments
angiogenesis and endothelial cell adhesion and migration, as well as in- [220,221]. Snake venoms contain biologically active peptides, which
creased microtubule dynamics. Therefore, this toxin is able to interfere are a rich source of possible novel antimicrobials agents [222]. Further-
in crucial mechanisms related to the process of tumorigenesis [83]. more, several well-characterized snake toxins, such as phospholipases
Some studies also proposed an antibleeding function for these toxins, A2, metalloproteases, L-amino acid oxidases and crotamine have shown
since some kunitz-type snake venom protease inhibitors are capable effects against many microorganisms [223,224].
830 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

Most of the studies regarding antimicrobial peptides from snake (PLB), the cleavage occurs at sn-1 and sn-2 positions [95]. PLBs were poor-
venoms report only the screening of crude venom antimicrobial ly reported so far since they are rarely found in snake venoms, in contrast
activity [225–228]. These peptides are mainly identified by mass spec- to phospholipases A2 (PLA2s) that are amongst the major venom
trometry following chromatographic isolation, although genomic se- components.
quencing and functional characterization by different antimicrobial There are several snake species that have PLB activity in their venom,
assays are also employed [46,47,104,220]. Among the previously de- such as Pseudechis porphyriacus, P. australis, N. naja, D. russelii, Cerastes
scribed snake venom antimicrobial peptides, it is possible to highlight vipera and Crotalus adamanteus [240–242]. PLBs were firstly reported
cathelicidins, viperidins, waprins, β-defensins, beyond other not yet in Pseudechis colletti venom as a dimeric protein of 16.5 kDa that pre-
classified molecules. sents high hemolytic activity in erythrocytes and cytotoxicity to rhabdo-
One of the major AMP families in mammals is composed by myosarcoma cells [96].
cathelicidin peptides that are cationic peptides with highly heteroge- The omics era not only facilitated but also fastened the discovery of
neous structures [229,230]. These peptides are composed of 12 to 80 novel PLBs. The proteomic analysis of the Australian snake Pseudechis
amino acid residues which are mainly folded in α-helical structures [230]. guttatus revealed two peptide sequences that shared identity to PLB.
The cathelicidin-BF, a cathelicidin-derived antimicrobial peptide, They were identified in one basic spot of 2D–PAGE and these two pep-
was the first cathelicidin peptide identified in reptiles (Bungarus tides covered 5% of the PLB sequence. The authors suggest that this
fasciatus snake) [48,49] and presents in vitro antimicrobial activity on basic toxin may contribute to the hemolysis observed in P. guttatus
bacterial and fungal species [48]. Other cathelicidin-derived peptides envenoming [97].
were identified in O. hannah (OH-CATH and its derivatives OH- Recently, proteins with PLB sequence identity were found in Micrurus
CATH30 and OH-CM6) [46,47,50,51], B. fasciatus (BF-CATH and its de- dumerilii snake venom proteome, although in small relative abundance
rivatives BF-30 and BF-15) [48,52–54], Hydrophis cyanocinctus (Hc- (b 0.1%). This venom is mainly composed of PLA2 and three-finger toxins
CATH) [55] and N. atra (NA-CATH) [46,56]. These peptides exhibit anti- [243]. Interestingly, one PLB was identified by mass spectrometry analysis
microbial activity against gram-positive and gram-negative bacterial in the venom of the Brazilian snake L. m. rhombeata. It was the first report
strains and fungi. Furthermore, it presents antitumor activity, as dem- of this class of protein in Lachesis genus [140].
onstrated against BF-CATH cells [48]. Although PLBs have been reported in some snake venom proteomes,
Another family of snake venom AMPs is the vipericidins (cathelicidin- the advent of transcriptome techniques has allowed the identification of
related peptides). This family comprises toxins from South American pit this class in different snake genus. One EST that encodes a PLB from a
vipers [57], such as crotalicidin (C. d. terrificus), lachesicidin (L. m. new lipase family was identified in the cDNA library of Drysdalia
rhombeata), batroxicidin (Bothrops atrox) and lutzicidin (Bothrops lutzi). coronoides venom gland. This EST comprises a signal peptide of 36
These venom-derived peptides showed in vitro antimicrobial activity amino acid residues and a mature protein of around 60 kDa. Moreover,
against several bacterial strains, including Streptococcus pyogenes, Staphy- the authors also found some peptides in D. coronoides venom proteome
lococcus aureus, Escherichia coli, Klebsiella pneumonia, Enterococcus faecalis, that correspond to this protein [244].
Acinetobacter baumannii and Pseudomonas aeruginosa [57]. Rokyta and colleagues [245] analyzed the transcriptome from C.
Omwaprin, a cationic antimicrobial peptide belonging to the waprin adamanteus venom gland and found one transcript that encodes a PLB
family, was isolated from O. microlepidotus snake venom [104]. This anti- precursor presenting a signal peptide of 36 amino acid residues and
microbial peptide is composed of 50 amino acid residues and contains a the mature PLB of 526 amino acid residues. It represents just 0.06% of
whey acid protein (WAP) domain in its structure (PDB ID. 3NGG). the whole transcriptome.
Omwaprin features in vitro antimicrobial activity against Bacillus Furthermore, the P. flavoviridis venom gland transcriptome revealed
megaterium and Staphylococcus warnei gram-positive bacterial strains that PLB comprises 0.14% of the transcripts, while in Ovophis okinavensis
[104]. On the other hand, the occurrence of waprins in the venoms from the percentage of sequences that encodes this protein was just 0.15%.
Thrasops jacksonii, Liophis poecilogyrus, Philodryas olfersii, Pseudoferania Some peptide sequences corresponding to PLB were also found in the
polylepis (formerly Enhydris polylepis and Rhabdophis tigrinus was report- proteome of these two snake venoms and they cover from 26.1 to
ed only through transcriptomic approaches [102]. Corrêa-Netto et al. 61.6% of the PLB sequence [45].
[103] reported that waprin-like peptides correspond to 0.2% of the total The venom-gland transcriptome of the coral snake Micrurus fulvius
toxins in the cDNA library of Micrurus altirostris venom gland. revealed one PLB sequence, representing 0.154% of total reads. This
The peptide Pep5Bj (1.37 kDa), isolated from Bothropoides jararaca venom is mostly composed of PLA2 (64.9% of toxins reads), which is re-
venom, was able to inhibit the growth of yeast (Candida albicans) and sponsible for the neurotoxic, cardiotoxic, myotoxic, anticoagulant and
phytopathogenic fungi (Fusarium oxysporum and Colletotrichum hemolytic effects of the envenoming [246].
lindemuthianum) cell cultures [231]. Another peptide of 2.49 kDa, Vonk and colleagues [16] analyzed the genome from the venom and
known as NAP (Naja Antibacterial Peptide), was isolated from Indian accessory glands of king cobra snake (O. hannah) and observed that they
cobra (N. naja) venom and showed in vitro potent antibacterial activity are mainly composed of three-finger toxins (venom) and lectins (acces-
against gram-negative and gram-positive bacterial strains [232]. sory gland). In addition to these findings, they also reported PLB tran-
Phylogenetic analysis enabled the identification of 13 β-defensin- scripts in these two glands. In contrast, PLB could not be found in the
like sequences from Bothrops spp. and Lachesis spp. snakes by PCR ap- venom proteome of this snake.
proach [233]. β-Defensins-like peptides have also been described in Although some PLBs have been identified in snake venoms by geno-
other snake venoms [234,235] and display distinct biological effects, in- mic, transcriptomic and proteomic approaches, they are poorly studied
cluding antimicrobial activity [236]. by traditional approach, mainly due to their low abundance in crude
These snake venom antimicrobial peptides can be valuable to the de- venoms [242]. Therefore the importance of omics approaches to the dis-
velopment of novel antimicrobial drugs [47,48,222,237], besides their closure of minor components such as PLB is reinforced and further
potential as biomarkers as well as diagnostic tools [238,239]. structural and functional studies regarding this neglected class of
venom protein are required.
3.7. Phospholipases B
3.8. Cobra venom factor
Phospholipases are enzymes that catalyze the cleavage of phospho-
lipids, releasing fatty acids and lysophospholipids. Depending on the Cobra venom factor (CVF) is a family of proteins that activates the
class of phospholipase, this reaction may take place in four different complement system. The CVF function was reported by [58]. However,
sites in membrane phospholipids. When catalyzed by phospholipase B before that, Flexner and Noguchi [247] had already demonstrated that
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 831

Naja genus venom produces hemolysis, which is dependent of the com- toxin. Recombinant ohanin may be helpful in the investigation of the
plement system, when incubated with mammalian erythrocytes and structure-function relationship of this class of toxins [101].
serum [248]. In 1971, it was shown that the venom of N. haje (Egyptian
cobra) selectively consumes the complement components C3 → C9 3.10. Dipeptidylpeptidase IV
[249]. Lately, several studies focused on CVF purification and complete
functional characterization [250–253]. Thereby, CVF has been studied Dipeptidylpeptidase IV (DPP IV) presents a wide distribution among
for more than a century. snake venoms [264]. It has been identified by proteome and tran-
CVF presents approximately 149 kDa and shares high identity de- scriptome approaches, for example, in Rhinocerophis alternatus (former-
gree with human complement component C3 (50% of identity, 69% of ly Bothrops alternatus), Gloydius brevicaudus, P. flavoviridis and O.
similarity) [62]. The C3 is the most important protein of the comple- okinavensis venoms [44,45,72] and its activity have been detected in a
ment system, being activated by all the three pathways: classical, lectin great variety of snake venoms [264,265]. The DPP IV G. brevicaudus
and alternative, besides their possible cross-reactivity [59]. The C3 acti- was partially isolated from this snake venom and the cDNA sequence
vation and the generation of its products lead to all biological functions was cloned from the venom gland cDNA library. Two DPP IV, named
of the complement system [254]. Since CVF is a C3-like protein, after the DPP IVa and DPP IVb, containing 751 amino acid residues were obtain-
cleavage of factor B (by factor D), it is turned into CVFBb, which acts as a ed. They contain a consensus sequence similar to serine proteases
C3-convertase. The continuous action of CVFBb is responsible for com- around the Ser616 and a putative catalytic triad (Ser616, Asp694 and
plement depletion and production of components of the terminal His76) is located at the C-terminus region. The mature protein present-
phase, including the cleavage of C5 and the generation of C5a. The C5a ed 116 kDa estimated by SDS-PAGE, and about 25% of their molecular
component is a powerful pro-inflammatory peptide. Along with C3a, mass is due to glycosylation. In fact, the sequences present 10 potential
C5a could be an important product for the snake since they increase N-glycosylation sites [265]. It is interesting that DPP IV is secreted into
the vascular permeability and blood flow in the bite site, allowing the the venom gland through exosome-like vesicles not through the com-
rapid diffusion of the toxic venom compounds into the blood stream. mon pathway by secretory granules and exocytosis [266].
Excellent detailed reviews can be found elsewhere [62,255]. The role of DPP IV during envenoming is still not clear, but these en-
So far, CVF is mainly present in Elapidae snake family and it has been zymes might contribute to alterations in the cardiovascular system
described in four different genera: Hemachatus, Naja, Ophiophagus and leading to hypotension by cleaving endogenous hypertensive factors.
Austrelaps [60–62,249]. Nevertheless, complement depletion caused Additionally, DPP IV may also present important roles in the immune
by CVF is most studied using the venoms of N. naja, Naja kaouthia and and neuroendocrine systems and in the glucose homeostasis [71].
N. atra, even being found in low amounts in these venoms. Laustsen et
al. [162] demonstrated, by proteomic analysis, that CVF accounts for 3.11. Aminopeptidases
0.1% of N. kaouthia whole venom, which is a very low percentage in
comparison to other components, such as PLA2 (13.5%), neurotoxins Aminopeptidases have been identified by “omics” techniques in the
(53.2%) and LAAO (0.4%). However, recently, CVF was also found in venoms of B. jararaca [43], O. okinavensis [45], P. flavoviridis [45] and G.
the Viperidae snake (Bothrops jararaca) using transcriptomic [256] brevicaudus [44]. Aminopeptidase activity was detected in the mice
and proteopeptidomic [257] analysis. envenoming by C. d. terrificus venom [267]. Furthermore, aminopepti-
Based on the low relative abundance of CVF in snake venoms and its dase A (APA) is released to the venom gland lumen through exosome-
important anti-complement effect, molecular biology emerges as a con- like vesicles, the same say as DPP IV [266].
venient approach to CVF production for therapeutic purposes. In this The rhiminopeptidase A is an APA isolated from Bitis rhinoceros
way, recombinant [62,258] and humanized [259] CVF are available which is an exo-metallopeptidase that removes acidic terminal residues
nowadays. from proteins and peptides at presence of calcium. No specificity for
acidic, basic or neutral pH was detected in the absence of calcium. Its
3.9. Vespryn molecular mass was estimated by SDS-PAGE being 150 kDa, and its se-
quence is composed of 951 amino acid residues [268].
Vespryn is a new family of venom proteins comprising ohanin from The aminopeptidases mammalian A and N act by different ways in
O. hannah [101], Thai cobrin (P82885) from N. kaouthia [260] and the renin-angiotensin system. Aminopeptidase N (APN) catalyzes the
ohanin-like proteins from N. n. atra [261], L. muta [98] and Tropidechis conversion of angiotensin III to angiotensin IV through the hydrolysis
carinatus [262]. However, only ohanin (the first family member) has of the N-terminal arginine from angiotensin III thus preventing hyper-
been well characterized so far. tension. In contrast, APA induces hypotension by degrading angiotensin
Ohanin was firstly purified in 2005 from the crude venom of O. II to angiotensin III through the hydrolysis of the N-terminal aspartic
hannah (king cobra). It presents a molecular mass of approximately acid since angiotensin III is less hypertensive that angiotensin II at pe-
12 kDa, with 107 amino acid residues and a single cysteine residue, ripheral sites [45]. Both angiotensins (II and III) also take part in the
which is a unique feature of this protein class. The amino acid sequence brain renin-angiotensin system, involved in the blood pressure control
of ohanin shares 93% of sequence identity with Thai cobrin isolated from and both present hypertensive action with affinity to the AT1 and AT2
N. kaouthia [101]. receptors [269].
Ohanin induces hypolocomotion and hypernociception in mice after Besides that, a leucine aminopeptidase was also identified and it
intraperitoneal injection in a dose-dependent manner. However, it is might be linked to the action of hemorrhagic snake venom
considered non-toxic since the analysis of gross pathology, after 24 h metalloproteases, other venom peptidases and the L-amino acid ox-
of its injection in mice, showed no signs of hemorrhage or necrosis in idase activity [45].
the brain, heart, lungs, kidneys, spleen and liver. Thus, the ohanin effects
are possibly important to slow down the mobility of the prey, assisting 3.12. Other minor snake venom toxins
in its capture [101].
The complete cDNA and genomic organization of ohanin was al- Table 2 summarizes the main minor snake venom toxins identified
ready reported. The ohanin-encoding cDNA (1558 bp) show that this only by “omic” approaches.
toxin is synthesized as a prepro-protein in the venom gland [263]. In
this context, due to ohanin low recovery from the venom (0.1%), heter- 3.12.1. Cytokine-like molecules
ologous expression of this protein has already been accomplished and Transcriptomic analyses have enabled the identification of cytokine-
the resulting recombinant form functionally resembles the native like molecules in the venoms of D. acutus, R. alternatus and B. gabonica.
832 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

Table 2
Snake venom proteins identified by “omic” techniques.

Protein class Snake venoms Approach used Possible function in the venom References
in the identification

Bungarus fasciatus, Naja atra,


Cathelicidins Transcriptome Antimicrobial activity [46,48]
Ophiophagus hannah
Deinagkistrodon acutus, Rhinocerophis Local inflammation and alterations in the
Cytokine-like molecules Transcriptome [72,270,271]
alternatus, Bitis gabonica vascular system
Crotalus adamanteus, Bothropoides
EF-hand protein Transcriptome and Proteome Unknown [276–278]
insularis
Insulin-like growth factor Ophiophagus hannah Transcriptome and Proteome Unknown [16,61,208]
Lysosomal acid lipase (LAL) Echis coloratus Transcriptome and Proteome Unknown [103,279]
Bothropoides jararaca, Bothrops fonsecai,
Metalloprotease inhibitors
Bothrops cotiara, Bothrops atrox, Bothrops
(poly-Gly- and poly-His- Proteome and Peptidome Inhibition of metalloprotease activity [77,209–213]
barnetti, Bothrops ayerbei, Bothriechis
poly-Gly-peptides and tripeptides)
supraciliaris, Bothrocophias campbelli
Neuroprotective protein Deinagkistrodon acutus Transcriptome Unknown [271]
Increased vascular permeability,
hypotension, inhibition of
Nucleosides Dendroaspis polylepis Proteome neurotransmitter release, sedation, [45,71,161–163]
locomotor depression, bradycardia,
influence in the toxin distribution
Transforming growth factor Deinagkistrodon acutus Transcriptome Unknown [271]
Sistrurus catenatus edwardsii, Echis
Renin-like aspartic protease Transcriptome and Proteome Local hypertension [280,281]
ocellatus
Transferrin-like proteins (TLP) Pseudechis australis Proteome Antimicrobial activity [282]
Deinagkistrodon acutus, Rhinocerophis
Tumor necrosis factor Transcriptome Local inflammation [72,271]
alternatus
Cerberus rynchops, Micrurus corallinus, Platelet aggregation induction and
Veficolins Transcriptome [103,274,275]
Philodryas olfersii interference in the fibrin formation

Their role in the envenoming remains unknown, but they possibly act in 4. Conclusions and perspectives
the vascular system, contributing to the local inflammatory process in-
duced by the venom [72,270,271]. Snake venoms are rich combinations of bioactive compounds that
act on specific biochemical and physiological targets of the prey or vic-
tim, causing an imbalance on their homeostasis. Thus, snake venoms are
3.12.2. Veficolins considered a source of harmful toxins, possessing a diversity of mole-
Ficolins are a group of proteins similar to complement-activating cules that mimic some natural regulatory components [1,284,285]. On
lectins present in different tissues. They show a collagen-like domain the other hand, snake venoms toxins are also considered promising
in the N-terminal and a fibrinogen-like domain in the C-terminal region molecules for the treatment and diagnosis of certain pathologies, as
and are related to the innate defense system [272–274]. Veficolins well as for research and biotechnological purpose [284 286–288]. Fig.
(venom ficolins) were identified for the first time as a venom toxin in 2 summarizes the main potential application of the minor venom pro-
the venom gland transcriptome of the colubrid snake Cerberus rynchops tein classes discussed herein. The most abundant toxins are easily iden-
[274], although Ching et al. [275] reported a veficolin from P. olfersii as a tified and purified from snake venoms by classical approaches. They
putative protein in its venom. The collagen-like domain of ryncolins have been extensively studied in the last decades and some of them
(the C. rynchops ficolins) may induce platelet aggregation while the di- have already been directly applied in the treatment of some disorders
merization of two C-terminal fibrinogen-like domains might interfere in or used as models for drug designs, such as serine proteases, disintegrins
fibrin formation by mimicking the C-terminal globular domain of fibrin- and bradykinin-potentiating peptides [288]. A classic example is Cap-
ogen. Veficolins were also identified in the transcriptome of Micrurus topril, which is an orally available peptidomimetic of a bradykinin
corallinus [103,274,275]. potentiating peptide from B. jararaca venom. This molecule compet-
itively inhibits the angiotensin-converting enzyme (ACE) and is ex-
tensively used worldwide in the treatment of human hypertension
3.12.3. EF-hand protein [289]. Currently, nine snake venom protein derivatives are commer-
EF-hand proteins present a conserved calcium-binding domain, cially available for therapeutic application or under clinical trials
composed of two alpha helixes linked by a loop (helix-loop-helix) usu- [288], but none of them are related to the protein classes approached
ally formed by 12 amino acid residues. This helix-loop-helix motif that in this review.
coordinates calcium-binding resembles the human hand thumb and Most of the minor snake venom proteins remained unidentified or
forefinger, which provided the name EF-hand for this motif. This do- poorly explored for many years, mainly due to the difficulties in
main is important to calcium-dependent processes, such as membrane obtaining them in sufficient amounts for a comprehensive study [21].
fusion and vesicular transport [276]. Calglandulin, an EF-hand protein Only with the advent of new highly sensitive methods of molecular
identified in the B. insularis venom gland, has been implicated in the se- identification, such as the “omic” approaches, this hidden world of
cretion of toxins from the venom gland cells [245,276,283]. On the other snake toxins could be brought to light. However, an in-depth investiga-
hand, reticulocalbin 2 EF-hand calcium-binding domain precursor is a tion of the biological effects and the potential for therapeutic and bio-
highly expressed nontoxin in the venom gland C. adamanteus [245, technological application of these toxins is still necessary.
277]. However, it is not yet clear if this molecule present toxic properties Heterologous protein expression and peptide synthesis represent al-
and potential role in the envenoming or if it presence in the venom pro- ternatives to produce toxins in both laboratory and industrial scale,
teome is just a result of the leakage of a housekeeping protein. which may enable the study of minor snake toxin classes and generate
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 833

Fig. 2. Minor snake venom protein classes and their main potential applications. The summary of the minor protein classes discussed herein are illustrated. Magenta: antimicrobial class.
Green: proteases inhibitors class. Yellow: growth factors class. Blue: other classes. The possible applications are numbered and indicated in the right panel, accordingly.

perspectives for their further investigation and application [21]. Further- effects of their homologous from nonvenomous tissues. However, stud-
more, we highlight the importance of continuing the search for novel ies related to the origin of these protein classes in the venom are recent
toxin scaffolds, classes and isoforms, in order to enrich our knowledge and still very scarce and inconclusive. The analysis of the differences in
on these natural libraries of pharmacologically active molecules. Improv- the expression rates of these proteins in the venom gland and in other
ing the understanding in snake proteins diversity is indispensable not tissues and the comparative characterization of venom proteins and
only to drug discovery, but also to better comprehend the pathophysiolo- their homologs may indicate whether protein may have evolved to dis-
gy of snake bites and, thereby, improving the still outdated antivenom play toxic effects or another role in these complexes biological. More-
therapy. over, investigations related to the gene structure and expression of
Minor snake venom protein classes may also provide important infor- venom proteins will assist the understanding of the evolutionary pro-
mation regarding the evolutionary processes that generated snake venom cesses that originated venom components and enlightening of the se-
repertories. Snake toxins are supposed to have been originated by genes lective forces that have been driven adaptation of snake venoms to
that were expressed both on venom gland and other tissues, followed different ecological niches and preys.
by tissue-specific enhanced expression in the venom gland and reduction
of expression or gene loss in nonvenomous-related tissues [256,290,291]. Transparency document
Evidences that support this hypothesis is the expression of various venom
homolog genes in nonvenom gland-related tissues of venomous and non- The Transparency document associated with this article can be
venomous snakes [291]. Subsequently, the main toxic and abundant found, in online version.
snake venom proteins probably suffered selective pressures and
underwent diversification by gene duplication that occurs via “birth and Acknowledgments
death” model [285], followed by mutations in key residues or domains
in the molecular surface [16,292] and/or gene recombination [29]. This The authors acknowledge the financial support from the Fundação
process typically results in multi-gene families, with conserved few disul- de Amparo à Pesquisa do Estado de São Paulo (FAPESP, São Paulo Re-
fide-rich molecular scaffolds, but encoding a variety of proteins exhibiting search Foundation, grant n. 2011/23236-4; scholarships to: GAW, n.
diverse biological and toxic activities [6,293,294]. The reason for generat- 2014/06170-8; KCFB, n. 2013/26619-7; JBF, n. 2015/16714-8; PYTS n.
ing toxin diversification in the venom is still not completely clarified. Nev- 2014/15644-3; CTC, n. 2013/26200-6; FAC, n. 2012/13590-8; MBP, n.
ertheless, it may represent a selective advantage to these animals since 2012/12954-6 and FGA, n. 2011/12317-3), Conselho Nacional de
different toxins may act on synergism, enhancing toxic effects, and may Desenvolvimento Científico e Tecnológico (CNPq, The National Council
extend molecular targets in preys [294,295]. for Scientific and Technological Development, 303689/2013-7; scholar-
On the other hand, genes encoding many low abundant snake ship to FAC n. 140949/2015-1), Coordenação de Aperfeiçoamento de
venom proteins were presumably devoid of selective pressures, mainly Pessoal de Nível Superior (CAPES, Coordination for the Improvement of
due to their small contribution in the prey capture [13]. Thus, the minor Higher Education Personnel, scholarship to GAW) (scholarship to ELPJ,
protein families discussed herein are expected to present low degree of ISO and IGF) and the Support Nucleus for Research on Animal Toxins
diversification and neofunctionalization, retaining most of the biological (NAP-TOXAN-USP, grant n. 12–125432.1.3).
834 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

References [29] N.L. Dowell, M.W. Giorgianni, V.A. Kassner, J.E. Selegue, E.E. Sanchez, S.B. Carroll,
The deep origin and recent loss of venom toxin genes in rattlesnakes, Curr. Biol.
26 (2016) 2434–2445.
[1] B.G. Fry, K. Roelants, D.E. Champagne, H. Scheib, J.D.A. Tyndall, G.F. King, T.J. [30] H.M. Kerkkamp, R.M. Kini, A.S. Pospelov, F.J. Vonk, C.V. Henkel, M.K. Richardson,
Nevalainen, J.A. Norman, R.J. Lewis, R.S. Norton, C. Renjifo, R.C.R. de la Vega, The Snake Genome Sequencing: Results and Future Prospects, Toxins (Basel) 8 (2016).
toxicogenomic multiverse: Convergent recruitment of proteins into animal [31] A. Zelanis, A.K. Tashima, Unraveling snake venom complexity with 'omics' ap-
venoms, Annu. Rev. Genomics Hum. Genet. 10 (2009) 483–511. proaches: challenges and perspectives, Toxicon 87 (2014) 131–134.
[2] N.R. Casewell, Venom evolution: Gene loss shapes phenotypic adaptation, Curr. [32] S.G. Conticello, Y. Gilad, N. Avidan, E. Ben-Asher, Z. Levy, M. Fainzilber, Mechanisms
Biol. 26 (2016) R849–R851. for evolving hypervariability: the case of conopeptides, Mol. Biol. Evol. 18 (2001)
[3] J.P. Chippaux, V. Williams, J. White, Snake venom variability: methods of study, re- 120–131.
sults and interpretation, Toxicon 29 (1991) 1279–1303. [33] S. Dutertre, A.H. Jin, Q. Kaas, A. Jones, P.F. Alewood, R.J. Lewis, Deep venomics re-
[4] J.J. Calvete, L. Sanz, P. Cid, P. de la Torre, M. Flores-Diaz, M.C. Dos Santos, A. Borges, veals the mechanism for expanded peptide diversity in cone snail venom, Mol.
A. Bremo, Y. Angulo, B. Lomonte, A. Alape-Giron, J.M. Gutierrez, Snake venomics of Cell. Proteomics 12 (2013) 312–329.
the Central American rattlesnake Crotalus simus and the South American Crotalus [34] D. Andrade-Silva, A. Zelanis, E.S. Kitano, I.L.M. Junqueira-de-Azevedo, M.S. Reis, A.S.
durissus complex points to neurotoxicity as an adaptive paedomorphic trend Lopes, S.M.T. Serrano, Proteomic and glycoproteomic profilings reveal that post-
along Crotalus dispersal in South America, J. Proteome Res. 9 (2010) 528–544. translational modifications of toxins contribute to venom phenotype in snakes, J.
[5] J.J. Calvete, L. Sanz, Y. Angulo, B. Lomonte, J.M. Gutierrez, Venoms, venomics, Proteome Res. 15 (2016) 2658–2675.
antivenomics, FEBS Lett. 583 (2009) 1736–1743. [35] J.R. Prashanth, R.J. Lewis, S. Dutertre, Towards an integrated venomics approach for
[6] D. Kordis, F. Gubensek, Molecular evolution of Bov-B LINEs in vertebrates, Gene accelerated conopeptide discovery, Toxicon 60 (2012) 470–477.
238 (1999) 171–178. [36] P. Favreau, L. Menin, S. Michalet, F. Perret, O. Cheneval, M. Stocklin, P. Bulet, R.
[7] H.I. Rosenberg, Histology, histochemistry, and emptying mechanism of the venom Stocklin, Mass spectrometry strategies for venom mapping and peptide sequenc-
glands of some elapid snakes, J. Morphol. 123 (1967) 133–155. ing from crude venoms: case applications with single arthropod specimen, Toxicon
[8] S.M. Carneiro, V.R. Pinto, C. Jared, L.A. Lula, F.P. Faria, A. Sesso, Morphometric stud- 47 (2006) 676–687.
ies on venom secretory cells from Bothrops jararacussu (Jararacuçu) before and [37] J.W. Fox, A brief review of the scientific history of several lesser-known snake
after venom extraction, Toxicon 29 (1991) 569–580. venom proteins: L-amino acid oxidases, hyaluronidases and phosphodiesterases,
[9] U. Oron, A. Bdolah, Regulation of protein synthesis in the venom gland of viperid Toxicon 62 (2013) 75–82.
snakes, J. Cell Biol. 56 (1973) 177–190. [38] J.J. Calvete, B. Lomonte, A bright future for integrative venomics, Toxicon 107
[10] E. Kochva, Oral glands of the Reptilia, Biology of the Reptilia, 8, 1978 43–161. (2015) 159–162.
[11] M.S. Luna, R.H. Valente, J. Perales, M.L. Vieira, N. Yamanouye, Activation of Bothrops [39] P.F. Campos, D. Andrade-Silva, A. Zelanis, A.F. Paes Leme, M.M. Rocha, M.C.
jararaca snake venom gland and venom production: A proteomic approach, J. Pro- Menezes, S.M. Serrano, I.L. Junqueira-de-Azevedo, Trends in the evolution of
teome 94 (2013) 460–472. snake toxins underscored by an integrative omics approach to profile the venom
[12] D.R. Rokyta, K.P. Wray, J.J. McGivem, M.J. Margres, The transcriptomic and proteo- of the colubrid Phalotris mertensi, Genome Biol Evol (2016).
mic basis for the evolution of a novel venom phenotype within the timber rattle- [40] X. Chen, X.D. Yu, M. Deng, H. Li, Q.Y. He, J.P. Liu, Purification and characterization of
snake (Crotalus horridus), Toxicon 98 (2015) 34–48. 5'-nucleotidase from Trimeresurus albolabris venom, Zool. Res. 29 (2008) 399–404.
[13] N.R. Casewell, S.C. Wagstaff, W. Wuster, D.A.N. Cook, F.M.S. Bolton, S.I. King, D. Pla, [41] M.L. Hart, D. Kohler, T. Eckle, D. Kloor, G.L. Stahl, H.K. Eltzschig, Direct treatment of
L. Sanz, J.J. Calvete, R.A. Harrison, Medically important differences in snake venom mouse or human blood with soluble 5'-nucleotidase inhibits platelet aggregation,
composition are dictated by distinct postgenomic mechanisms, Proc. Natl. Acad. Arterioscler. Thromb. Vasc. Biol. 28 (2008) 1477–1483.
Sci. U. S. A. 111 (2014) 9205–9210. [42] C. Ouyang, T.F. Huang, Inhibition of platelet aggregation by 5′-nucleotidase purified
[14] J. Durban, A. Perez, L. Sanz, A. Gomez, F. Bonilla, S. Rodriguez, D. Chacon, M. Sasa, Y. from Trimeresurus gramineus snake venom, Toxicon 21 (1983) 491–501.
Angulo, J.M. Gutierrez, J.J. Calvete, Integrated "omics" profiling indicates that [43] G.S. Dias, E.S. Kitano, A.H. Pagotto, S.S. Sant'anna, M.M. Rocha, A. Zelanis, S.M.
miRNAs are modulators of the ontogenetic venom composition shift in the Central Serrano, Individual variability in the venom proteome of juvenile Bothrops jararaca
American rattlesnake, Crotalus simus simus, BMC Genomics 14 (2013). specimens, J. Proteome Res. 12 (2013) 4585–4598.
[15] J. Boldrini-Franca, C. Correa-Netto, M.M.S. Silva, R.S. Rodrigues, P. De La Torre, A. [44] J.F. Gao, Y.F. Qu, X.Q. Zhang, Y. He, X. Ji, Neonate-to-adult transition of snake
Perez, A.M. Soares, R.B. Zingali, R.A. Nogueira, V.M. Rodrigues, L. Sanz, J.J. Calvete, venomics in the short-tailed pit viper, Gloydius brevicaudus, J. Proteome 84
Snake venomics and antivenomics of Crotalus durissus subspecies from Brazil: As- (2013) 148–157.
sessment of geographic variation and its implication on snakebite management, J. [45] S.D. Aird, Y. Watanabe, A. Villar-Briones, M.C. Roy, K. Terada, A.S. Mikheyev, Quan-
Proteome 73 (2010) 1758–1776. titative high-throughput profiling of snake venom gland transcriptomes and
[16] F.J. Vonk, N.R. Casewell, C.V. Henkel, A.M. Heimberg, H.J. Jansen, R.J. McCleary, H.M. proteomes (Ovophis okinavensis and Protobothrops flavoviridis), BMC Genomics
Kerkkamp, R.A. Vos, I. Guerreiro, J.J. Calvete, W. Wüster, A.E. Woods, J.M. Logan, 14 (2013) 790.
R.A. Harrison, T.A. Castoe, A.P. de Koning, D.D. Pollock, M. Yandell, D. Calderon, C. [46] H. Zhao, T.X. Gan, X.D. Liu, Y. Jin, W.H. Lee, J.H. Shen, Y. Zhang, Identification and
Renjifo, R.B. Currier, D. Salgado, D. Pla, L. Sanz, A.S. Hyder, J.M. Ribeiro, J.W. characterization of novel reptile cathelicidins from elapid snakes, Peptides 29
Arntzen, G.E. van den Thillart, M. Boetzer, W. Pirovano, R.P. Dirks, H.P. Spaink, D. (2008) 1685–1691.
Duboule, E. McGlinn, R.M. Kini, M.K. Richardson, The king cobra genome reveals [47] S.A. Li, W.H. Lee, Y. Zhang, Efficacy of OH-CATH30 and its analogs against drug-re-
dynamic gene evolution and adaptation in the snake venom system, Proc. Natl. sistant bacteria in vitro and in mouse models, Antimicrob. Agents Chemother. 56
Acad. Sci. U. S. A. 110 (2013) 20651–20656. (2012) 3309–3317.
[17] M.D. McCue, Cost of producing venom in three North American pitviper species, [48] Y. Wang, J. Hong, X. Liu, H. Yang, R. Liu, J. Wu, A. Wang, D. Lin, R. Lai, Snake
Copeia (2006) 818–825. cathelicidin from Bungarus fasciatus is a potent peptide antibiotics, PLoS One 3
[18] D. Rotenberg, E.S. Bamberger, E. Kochva, Studies on ribonucleic acid synthesis in (2008) e3217.
the venom glands of Vipera palaestinae (Ophidia, Reptilia), Biochem. J. 121 [49] Y. Wang, Z. Zhang, L. Chen, H. Guang, Z. Li, H. Yang, J. Li, D. You, H. Yu, R. Lai,
(1971) 609–612. Cathelicidin-BF, a snake cathelicidin-derived antimicrobial peptide, could be an ex-
[19] W.K. Hayes, The snake venom-metering controversy: levels of analysis, assump- cellent therapeutic agent for Acne vulgaris, PLoS One 6 (2011) e22120.
tions, and evidence, The biology of rattlesnakes 2008, pp. 191–220. [50] Y. Zhang, H. Zhao, G.Y. Yu, X.D. Liu, J.H. Shen, W.H. Lee, Y. Zhang, Structure-function
[20] D. Morgenstern, G.F. King, The venom optimization hypothesis revisited, Toxicon relationship of king cobra cathelicidin, Peptides 31 (2010) 1488–1493.
63 (2013) 120–128. [51] B.Y. Zhang, S.M. Li, Z.H. Gao, J.H. Shen, Protective effects of snake venom antimicro-
[21] J. Boldrini-Franca, R. Santos Rodrigues, L.K. Santos-Silva, D.L. de Souza, M.S. Gomes, bial peptide OH-CATH on E. coli induced rabbit urinary tract infection models, Zool.
C.T. Cologna, E. de Pauw, L. Quinton, F. Henrique-Silva, V. de Melo Rodrigues, E.C. Res. 34 (2013) 25–32.
Arantes, Expression of a new serine protease from Crotalus durissus collilineatus [52] W. Chen, B. Yang, H. Zhou, L. Sun, J. Dou, H. Qian, W. Huang, Y. Mei, J. Han, Struc-
venom in Pichia pastoris and functional comparison with the native enzyme, ture-activity relationships of a snake cathelicidin-related peptide, BF-15, Peptides
Appl. Microbiol. Biotechnol. 99 (2015) 9971–9986. 32 (2011) 2497–2503.
[22] J.J. Calvete, Venomics, what else? Toxicon 60 (2012) 427–433. [53] H. Wang, M. Ke, Y. Tian, J. Wang, B. Li, Y. Wang, J. Dou, C. Zhou, BF-30 selectively
[23] Q. Kaas, D.J. Craik, Bioinformatics-aided venomics, Toxins (Basel) 7 (2015) inhibits melanoma cell proliferation via cytoplasmic membrane permeabilization
2159–2187. and DNA-binding in vitro and in B16F10-bearing mice, Eur. J. Pharmacol. 707
[24] R.K. Brahma, R.J. McCleary, R.M. Kini, R. Doley, Venom gland transcriptomics for (2013) 1–10.
identifying, cataloging, and characterizing venom proteins in snakes, Toxicon 93 [54] H. Zhou, J. Dou, J. Wang, L. Chen, H. Wang, W. Zhou, Y. Li, C. Zhou, The antibacterial
(2015) 1–10. activity of BF-30 in vitro and in infected burned rats is through interference with
[25] K. Sunagar, D. Morgenstern, A.M. Reitzel, Y. Moran, Ecological venomics: How ge- cytoplasmic membrane integrity, Peptides 32 (2011) 1131–1138.
nomics, transcriptomics and proteomics can shed new light on the ecology and [55] L. Wei, J. Gao, S. Zhang, S. Wu, Z. Xie, G. Ling, Y.Q. Kuang, Y. Yang, H. Yu, Y. Wang,
evolution of venom, J. Proteome 135 (2016) 62–72. Identification and characterization of the first cathelicidin from sea snakes with
[26] A. Bazaa, N. Marrakchi, M. El Ayeb, L. Sanz, J.J. Calvete, Snake venomics: compara- potent antimicrobial and anti-inflammatory activity and special mechanism, J.
tive analysis of the venom proteomes of the Tunisian snakes Cerastes cerastes, Biol. Chem. 290 (2015) 16633–16652.
Cerastes vipera and Macrovipera lebetina, Proteomics 5 (2005) 4223–4235. [56] R.J. Blower, S.M. Barksdale, M.L. van Hoek, Snake cathelicidin NA-CATH and smaller
[27] P. Juarez, L. Sanz, J.J. Calvete, Snake venomics: characterization of protein families helical antimicrobial peptides are effective against Burkholderia thailandensis, PLoS
in Sistrurus barbouri venom by cysteine mapping, N-terminal sequencing, and tan- Negl. Trop. Dis. 9 (2015) e0003862.
dem mass spectrometry analysis, Proteomics 4 (2004) 327–338. [57] C.B. Falcao, B.G. de La Torre, C. Perez-Peinado, A.E. Barron, D. Andreu, G. Radis-
[28] A. Menez, R. Stocklin, D. Mebs, 'Venomics' or : The venomous systems genome pro- Baptista, Vipericidins: a novel family of cathelicidin-related peptides from the
ject, Toxicon 47 (2006) 255–259. venom gland of South American pit vipers, Amino Acids 46 (2014) 2561–2571.
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 835

[58] C.B. Ewing, The action of rattlesnake venom upon the bactericidal power of the [87] J.C. Perez, E.E. Sanchez, Natural protease inhibitors to hemorrhagins in snake
blood serum, Boston Med. Surg. J. 130 (1894) 487–490. venoms and their potential use in medicine, Toxicon 37 (1999) 703–728.
[59] A. Laich, R.B. Sim, Complement C4bC2 complex formation: an investigation by sur- [88] R.J. Mannion, M. Costigan, I. Decosterd, F. Amaya, Q.P. Ma, J.C. Holstege, R.R. Ji, A.
face plasmon resonance, Biochim. Biophys. Acta 1544 (2001) 96–112. Acheson, R.M. Lindsay, G.A. Wilkinson, C.J. Woolf, Neurotrophins: peripherally
[60] S. Rehana, R. Manjunatha Kini, Molecular isoforms of cobra venom factor-like pro- and centrally acting modulators of tactile stimulus-induced inflammatory pain hy-
teins in the venom of Austrelaps superbus, Toxicon 50 (2007) 32–52. persensitivity, Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 9385–9390.
[61] C.H. Tan, K.Y. Tan, S.Y. Fung, N.H. Tan, Venom-gland transcriptome and venom pro- [89] K. Sunagar, B.G. Fry, T.N. Jackson, N.R. Casewell, E.A. Undheim, N. Vidal, S.A. Ali, G.F.
teome of the Malaysian king cobra (Ophiophagus hannah), BMC Genomics 16 King, K. Vasudevan, V. Vasconcelos, A. Antunes, Molecular evolution of vertebrate
(2015) 687. neurotrophins: co-option of the highly conserved nerve growth factor gene into
[62] C.W. Vogel, D.C. Fritzinger, Cobra venom factor: Structure, function, and humaniza- the advanced snake venom arsenal, PLoS One 8 (2013) e81827.
tion for therapeutic complement depletion, Toxicon 56 (2010) 1198–1222. [90] T. Kostiza, C.A. Dahinden, S. Rihs, U. Otten, J. Meier, Nerve growth factor from the
[63] Q. Xie, N. Tang, Y. Lin, X. Wang, X. Lin, J. Lin, Recombinant adenovirus snake venom venom of the Chinese cobra Naja naja atra: purification and description of non-
cystatin inhibits the growth, invasion, and metastasis of B16F10 cells in vitro and in neuronal activities, Toxicon 33 (1995) 1249–1261.
vivo, Melanoma Res. 23 (2013) 444–451. [91] B.L. Dhananjaya, D.S. CJ, An overview on nucleases (DNase, RNase, and phosphodi-
[64] Q. Xie, N. Tang, R. Wan, Y. Qi, X. Lin, J. Lin, Recombinant snake venom cystatin in- esterase) in snake venoms, Biochemistry (Mosc) 75 (2010) 1–6.
hibits the growth, invasion and metastasis of B16F10 cells and MHCC97H cells in [92] F.E. Russell, F.W. Buess, J. Strassbe, Zootoxicological properties of venom phospho-
vitro and in vivo, Toxicon 57 (2011) 704–711. diesterase, Toxicon 1 (1963) 99–108.
[65] Q. Xie, N. Tang, R. Wan, Y. Qi, X. Lin, J. Lin, Recombinant snake venom cystatin in- [93] M.L. Santoro, T.S. Vaquero, A.F. Paes Leme, S.M. Serrano, NPP-BJ, a nucleotide
hibits tumor angiogenesis in vitro and in vivo associated with downregulation of pyrophosphatase/phosphodiesterase from Bothrops jararaca snake venom, inhibits
VEGF-A165, Flt-1 and bFGF, Anti Cancer Agents Med. Chem. 13 (2013) 663–671. platelet aggregation, Toxicon 54 (2009) 499–512.
[66] Y. Yamazaki, T. Morita, Structure and function of snake venom cysteine-rich secre- [94] S. Uzawa, Über die phosphomonoesterase und die phosphodiesterase, J. Biochem.
tory proteins, Toxicon 44 (2004) 227–231. 15 (1932) 1–10.
[67] Y. Yamazaki, H. Koike, Y. Sugiyama, K. Motoyoshi, T. Wada, S. Hishinuma, M. Mita, [95] A. Aloulou, Y.B. Ali, S. Bezzine, Y. Gargouri, M.H. Gelb, Phospholipases: an overview,
T. Morita, Cloning and characterization of novel snake venom proteins that block Methods Mol. Biol. 861 (2012) 63–85.
smooth muscle contraction, Eur. J. Biochem. 269 (2002) 2708–2715. [96] A.W. Bernheimer, R. Linder, S.A. Weinstein, K.S. Kim, Isolation and characterization
[68] R.L. Brown, L.L. Lynch, T.L. Haley, R. Arsanjani, Pseudechetoxin binds to the pore of a phospholipase B from venom of Collett's snake, Pseudechis colletti, Toxicon 25
turret of cyclic nucleotide-gated ion channels, J. Gen. Physiol. 122 (2003) 749–760. (1987) 547–554.
[69] J. Wang, B. Shen, M. Guo, X. Lou, Y. Duan, X.P. Cheng, M. Teng, L. Niu, Q. Liu, Q. [97] V.L. Viala, D. Hildebrand, M. Trusch, R.K. Arni, D.C. Pimenta, H. Schluter, C. Betzel,
Huang, Q. Hao, Blocking effect and crystal structure of natrin toxin, a cysteine- P.J. Spencer, Pseudechis guttatus venom proteome: Insights into evolution and
rich secretory protein from Naja atra venom that targets the BKCa channel, Bio- toxin clustering, J. Proteome 110 (2014) 32–44.
chemistry 44 (2005) 10145–10152. [98] I.L. Junqueira-de-Azevedo, A.T. Ching, E. Carvalho, F. Faria, M.Y. Nishiyama Jr., P.L.
[70] F. Wang, H. Li, M.N. Liu, H. Song, H.M. Han, Q.L. Wang, C.C. Yin, Y.C. Zhou, Z. Qi, Y.Y. Ho, M.R. Diniz, Lachesis muta (Viperidae) cDNAs reveal diverging pit viper mole-
Shu, Z.J. Lin, T. Jiang, Structural and functional analysis of natrin, a venom protein cules and scaffolds typical of cobra (Elapidae) venoms: implications for snake
that targets various ion channels, Biochem. Biophys. Res. Commun. 351 (2006) toxin repertoire evolution, Genetics 173 (2006) 877–889.
443–448. [99] Z.K. Otrock, J.A. Makarem, A.I. Shamseddine, Vascular endothelial growth factor
[71] S.D. Aird, Ophidian envenomation strategies and the role of purines, Toxicon 40 family of ligands and receptors: review, Blood Cells Mol. Dis. 38 (2007) 258–268.
(2002) 335–393. [100] Y. Yamazaki, K. Takani, H. Atoda, T. Morita, Snake venom vascular endothelial
[72] K.C. Cardoso, M.J. Da Silva, G.G. Costa, T.T. Torres, L.E. Del Bem, R.O. Vidal, M. growth factors (VEGFs) exhibit potent activity through their specific recognition
Menossi, S. Hyslop, A transcriptomic analysis of gene expression in the venom of KDR (VEGF receptor 2), J. Biol. Chem. 278 (2003) 51985–51988.
gland of the snake Bothrops alternatus (urutu), BMC Genomics 11 (2010) 605. [101] Y.F. Pung, P.T. Wong, P.P. Kumar, W.C. Hodgson, R.M. Kini, Ohanin, a novel protein
[73] K.C.F. Bordon, G.A. Wiezel, F.G. Amorim, E.C. Arantes, Arthropod venom Hyaluron- from king cobra venom, induces hypolocomotion and hyperalgesia in mice, J. Biol.
idases: biochemical properties and potential applications in medicine and biotech- Chem. 280 (2005) 13137–13147.
nology, J Venom Anim Toxins incl Trop Dis 21 (2015). [102] B.G. Fry, H. Scheib, L. van der Weerd, B. Young, J. McNaughtan, S.F. Ramjan, N. Vidal,
[74] S.P. Mackessy, Handbook of venoms and toxins of reptiles, Taylor & Francis, 2010. R.E. Poelmann, J.A. Norman, Evolution of an arsenal: structural and functional di-
[75] S. Pukrittayakamee, D.A. Warrell, V. Desakorn, A.J. Mcmichael, N.J. White, D. versification of the venom system in the advanced snakes (Caenophidia), Mol.
Bunnag, The hyaluronidase activities of some Southeast Asian snake venoms, Cell. Proteomics 7 (2008) 215–246.
Toxicon 26 (1988) 629–637. [103] C. Corrêa-Netto, I.e.L. Junqueira-de-Azevedo, D.A. Silva, P.L. Ho, M. Leitão-de-
[76] B. Lomonte, J. Escolano, J. Fernandez, L. Sanz, Y. Angulo, J.M. Gutierrez, J.J. Calvete, Araújo, M.L. Alves, L. Sanz, D. Foguel, R.B. Zingali, J.J. Calvete, Snake venomics and
Snake venomics and antivenomics of the arboreal neotropical pitvipers Bothriechis venom gland transcriptomic analysis of Brazilian coral snakes, Micrurus altirostris
lateralis and Bothriechis schlegelii, J. Proteome Res. 7 (2008) 2445–2457. and M. corallinus, J. Proteome 74 (2011) (1795-1809).
[77] B. Lomonte, W.C. Tsai, F. Bonilla, A. Solorzano, G. Solano, Y. Angulo, J.M. Gutierrez, [104] D.G. Nair, B.G. Fry, P. Alewood, P.P. Kumar, R.M. Kini, Antimicrobial activity of
J.J. Calvete, Snake venomics and toxicological profiling of the arboreal pitviper omwaprin, a new member of the waprin family of snake venom proteins,
Bothriechis supraciliaris from Costa Rica, Toxicon 59 (2012) 592–599. Biochem. J. 402 (2007) 93–104.
[78] W. Chen, L.P. Carvalho, M.Y. Chan, R.M. Kini, T.S. Kang, Fasxiator, a novel factor XIa [105] R.J. Wordinger, A.F. Clark, Growth factors and neurotrophic factors as targets, in: T.
inhibitor from snake venom, and its site-specific mutagenesis to improve potency Yorio, A.F. Clark, M.B. Wax (Eds.), Ocular therapeutics: Eye on new discoveries, Ac-
and selectivity, J. Thromb. Haemost. 13 (2015) 248–261. ademic Press 2008, pp. 87–116.
[79] S.T.H. Earl, R. Richards, L.A. Johnson, S. Flight, S. Anderson, A. Liao, J. de Jersey, P.P. [106] L. Aloe, Rita Levi-Montalcini: the discovery of nerve growth factor and modern
Masci, M.F. Lavin, Identification and characterisation of kunitz-type plasma kalli- neurobiology, Trends Cell Biol. 14 (2004) 395–399.
krein inhibitors unique to Oxyuranus sp. snake venoms, Biochimie 94 (2012) [107] S. Cohen, R. Levi-Montalcini, Purification and properties of a nerve growth-pro-
365–373. moting factor isolated from mouse sarcoma-180, Cancer Res. 17 (1957) 15–20.
[80] C.T. Guo, S. McClean, C. Shaw, P.F. Rao, M.Y. Ye, A.J. Bjourson, Trypsin and chymo- [108] S. Cohen, R. Levi-Montalcini, A nerve growth-stimulating factor isolated from
trypsin inhibitor peptides from the venom of Chinese Daboia russellii siamensis, snake venom, Proc. Natl. Acad. Sci. U. S. A. 42 (1956) 571-&.
Toxicon 63 (2013) 154–164. [109] S. Cohen, The stimulation of epidermal proliferation by a specific protein (EGF),
[81] A.K. Mukherjee, S.P. Mackessy, S. Dutta, Characterization of a kunitz-type protease Dev. Biol. 12 (1965) 394–407.
inhibitor peptide (Rusvikunin) purified from Daboia russelii russelii venom, Int. J. [110] R. Ross, J. Glomset, B. Kariya, L. Harker, A platelet-dependent serum factor that
Biol. Macromol. 67 (2014) 154–162. stimulates the proliferation of arterial smooth muscle cells in vitro, Proc. Natl.
[82] M. Morjen, O. Kallech-Ziri, A. Bazaa, H. Othman, K. Mabrouk, R. Zouari-Kessentini, Acad. Sci. U. S. A. 71 (1974) 1207–1210.
L. Sanz, J.J. Calvete, N. Srairi-Abid, M. El Ayeb, J. Luis, N. Marrakchi, PIVL, a new ser- [111] T. Kostiza, J. Meier, Nerve growth factors from snake venoms: chemical properties,
ine protease inhibitor from Macrovipera lebetina transmediterranea venom, im- mode of action and biological significance, Toxicon 34 (1996) 787–806.
pairs motility of human glioblastoma cells, Matrix Biol. 32 (2013) 52–62. [112] M.F. Lavin, S. Earl, G. Birrell, L. St. Pierre, L. Guddat, J. Jersey, P. Masci, Snake venom
[83] M. Morjen, S. Honore, A. Bazaa, Z. Abdelkafi-Koubaa, A. Ellafi, K. Mabrouk, H. nerve growth factors, in: S.P. Mackessy (Ed.), Handbook of venoms and toxins of
Kovacic, M. El Ayeb, N. Marrakchi, J. Luis, PIVL, a snake venom kunitz-type serine reptiles, CRC Press/Taylor & Francis 2010, pp. 377–391.
protease inhibitor, inhibits in vitro and in vivo angiogenesis, Microvasc. Res. 95 [113] R.H. Angeletti, Nerve growth factor from cobra venom, Proc. Natl. Acad. Sci. U. S. A.
(2014) 149–156. 65 (1970) 668–674.
[84] S.M. Flight, L.A. Johnson, Q.S. Du, R.L. Warner, M. Trabi, P.J. Gaffney, M.F. Lavin, J. de [114] S.T. Earl, G.W. Birrell, T.P. Wallis, L.D. St Pierre, P.P. Masci, J. de Jersey, J.J. Gorman,
Jersey, P.P. Masci, Textilinin-1, an alternative anti-bleeding agent to aprotinin: Im- M.F. Lavin, Post-translational modification accounts for the presence of varied
portance of plasmin inhibition in controlling blood loss, Br. J. Haematol. 145 (2009) forms of nerve growth factor in Australian elapid snake venoms, Proteomics 6
207–211. (2006) 6554–6565.
[85] P.P. Masci, A.N. Whitaker, L.G. Sparrow, J. de Jersey, D.J. Winzor, D.J. Watters, M.F. [115] D. Koh, A. Armugam, K. Jeyaseelan, Sputa nerve growth factor forms a preferable
Lavin, P.J. Gaffney, Textilinins from Pseudonaja textilis textilis. Characterization of substitute to mouse 7S-beta nerve growth factor, Biochem. J. 383 (2004) 149–158.
two plasmin inhibitors that reduce bleeding in an animal model, Blood Coagul. Fi- [116] F.L. Pearce, B.E. Banks, D.V. Banthorpe, A.R. Berry, H.S. Davies, C.A. Vernon, The iso-
brinolysis 11 (2000) 385–393. lation and characterization of nerve-growth factor from the venom of Vipera
[86] E.K. Millers, M. Trabi, P.P. Masci, M.F. Lavin, J. de Jersey, L.W. Guddat, Crystal struc- russelli, Eur. J. Biochem. 29 (1972) 417–425.
ture of textilinin-1, a kunitz-type serine protease inhibitor from the venom of the [117] R.A. Hogue-Angeletti, W.A. Frazier, J.W. Jacobs, H.D. Niall, R.A. Bradshaw, Purifica-
Australian common brown snake (Pseudonaja textilis), FEBS J. 276 (2009) tion, characterization, and partial amino acid sequence of nerve growth factor
3163–3175. from cobra venom, Biochemistry 15 (1976) 26–34.
836 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

[118] J. Siigur, U. Arumae, T. Neuman, E. Siigur, M. Saarma, Monoclonal antibody immu- [145] G.P. Queiroz, L.A. Pessoa, F.C. Portaro, F. Furtado Mde, D.V. Tambourgi, Interspecific
noaffinity chromatography of the nerve growth factor from snake venoms, Comp. variation in venom composition and toxicity of Brazilian snakes from Bothrops
Biochem. Physiol. B Biochem. Mol. Biol. 87 (1987) 329–334. genus, Toxicon 52 (2008) 842–851.
[119] L. Sanz, H.L. Gibbs, S.P. Mackessy, J.J. Calvete, Venom proteomes of closely related [146] C. Guerra-Duarte, J. Lopes-Peixoto, B.R. Fonseca-de-Souza, S. Stransky, D. Oliveira,
Sistrurus rattlesnakes with divergent diets, J. Proteome Res. 5 (2006) 2098–2112. F.S. Schneider, L. Lopes-de-Souza, C. Bonilla, W. Silva, B. Tintaya, A. Yarleque, C.
[120] L.Y. Guo, J.F. Zhu, X.F. Wu, Y.C. Zhou, Cloning of a cDNA encoding a nerve growth Chavez-Olortegui, Partial in vitro analysis of toxic and antigenic activities of eleven
factor precursor from the Agkistrodon halys Pallas, Toxicon 37 (1999) 465–470. Peruvian pitviper snake venoms, Toxicon 108 (2015) 84–96.
[121] S. Kashima, A.M. Soares, P.G. Roberto, J.O. Pereira, S. Astolfi-Filho, A.O. Cintra, M.R. [147] R. Shashidharamurthy, D.K. Jagadeesha, K.S. Girish, K. Kemparaju, Variations in bio-
Fontes, J.R. Giglio, S. de Castro Franca, cDNA sequence and molecular modeling of a chemical and pharmacological properties of Indian cobra (Naja naja naja) venom
nerve growth factor from Bothrops jararacussu venomous gland, Biochimie 84 due to geographical distribution, Mol. Cell. Biochem. 229 (2002) 93–101.
(2002) 675–680. [148] R. Shashidharamurthy, K. Kemparaju, Region-specific neutralization of Indian
[122] A.V. Osipov, T.I. Terpinskaya, E.V. Kryukova, V.S. Ulaschik, L.V. Paulovets, E.A. Petrova, cobra (Naja naja) venom by polyclonal antibody raised against the eastern region-
E.V. Blagun, V.G. Starkov, Y.N. Utkin, Nerve growth factor from cobra venom inhibits al venom: A comparative study of the venoms from three different geographical
the growth of Ehrlich tumor in mice, Toxins (Basel) 6 (2014) 784–795. distributions, Int. Immunopharmacol. 7 (2007) 61–69.
[123] Y. Yamazaki, T. Morita, Molecular and functional diversity of vascular endothelial [149] S. Pukrittayakamee, A. Nontprasert, N.J. White, D.A. Warrell, D. Bunnag, Character-
growth factors, Mol. Divers. 10 (2006) 515–527. ization of a monoclonal antibody that neutralizes the hyaluronidase activity of
[124] J. Boldrini-Franca, R.S. Rodrigues, F.P. Fonseca, D.L. Menaldo, F.B. Ferreira, F. Russell's viper venom, Southeast Asian J. Trop. Med. Public Health 21 (1990)
Henrique-Silva, A.M. Soares, A. Hamaguchi, V.M. Rodrigues, A.R. Otaviano, M.I. 231–237.
Homsi-Brandeburgo, Crotalus durissus collilineatus venom gland transcriptome: [150] K.S. Girish, K. Kemparaju, Inhibition of Naja naja venom hyaluronidase by plant-de-
analysis of gene expression profile, Biochimie 91 (2009) 586–595. rived bioactive components and polysaccharides, Biochemistry (Mosc) 70 (2005)
[125] L. Junqueira-de-Azevedo Ide, P.L. Ho, A survey of gene expression and diversity in 948–952.
the venom glands of the pitviper snake Bothrops insularis through the generation of [151] S.A. Khanum, S.K. Murari, B.S. Vishwanth, S. Shashikanth, Synthesis of benzoyl phe-
expressed sequence tags (ESTs), Gene 299 (2002) 279–291. nyl benzoates as effective inhibitors for phospholipase A2 and hyaluronidase en-
[126] R.S. Rodrigues, J. Boldrini-Franca, F.P. Fonseca, P. de la Torre, F. Henrique-Silva, L. zymes, Bioorg. Med. Chem. Lett. 15 (2005) 4100–4104.
Sanz, J.J. Calvete, V.M. Rodrigues, Combined snake venomics and venom gland [152] K. Sunitha, P. Suresh, M.S. Santhosh, M. Hemshekhar, R.M. Thushara, G.K. Marathe,
transcriptomic analysis of Bothropoides pauloensis, J. Proteome 75 (2012) C. Thirunavukkarasu, K. Kemparaju, M.S. Kumar, K.S. Girish, Inhibition of hyaluron-
2707–2720. idase by N-acetyl cysteine and glutathione: role of thiol group in hyaluronan pro-
[127] I.S. Sheen, K.S. Jeng, S.C. Shih, C.R. Kao, W.H. Chang, H.Y. Wang, P.C. Wang, T.E. tection, Int. J. Biol. Macromol. 55 (2013) 39–46.
Wang, L.R. Shyung, C.Z. Chen, Clinical significance of the expression of isoform [153] M. Molander, L. Nielsen, S. Sogaard, D. Staerk, N. Ronsted, D. Diallo, K.Z. Chifundera,
165 vascular endothelial growth factor mRNA in noncancerous liver remnants of J. van Staden, A.K. Jager, Hyaluronidase, phospholipase A2 and protease inhibitory
patients with hepatocellular carcinoma, World J. Gastroenterol. 11 (2005) activity of plants used in traditional treatment of snakebite-induced tissue necrosis
187–192. in Mali, DR Congo and South Africa, J. Ethnopharmacol. 157 (2014) 171–180.
[128] Y. Komori, T. Nikai, K. Taniguchi, K. Masuda, H. Sugihara, Vascular endothelial [154] Y. Liu, D. Staerk, M.N. Nielsen, N. Nyberg, A.K. Jager, High-resolution hyaluronidase
growth factor VEGF-like heparin-binding protein from the venom of Vipera aspis inhibition profiling combined with HPLC-HRMS-SPE-NMR for identification of
aspis (Aspic viper), Biochemistry 38 (1999) 11796–11803. anti-necrosis constituents in Chinese plants used to treat snakebite, Phytochemis-
[129] I.L. Junqueira de Azevedo, S.H. Farsky, M.L. Oliveira, P.L. Ho, Molecular cloning and try 119 (2015) 62–69.
expression of a functional snake venom vascular endothelium growth factor [155] A.N. Nanjaraj Urs, M. Yariswamy, V. Joshi, K.N. Suvilesh, M.S. Sumanth, D. Das, A.
(VEGF) from the Bothrops insularis pit viper. A new member of the VEGF family Nataraju, B.S. Vishwanath, Local and systemic toxicity of Echis carinatus venom:
of proteins, J. Biol. Chem. 276 (2001) 39836–39842. neutralization by Cassia auriculata L. leaf methanol extract, J. Nat. Med. 69 (2015)
[130] Y. Yamazaki, Y. Tokunaga, K. Takani, T. Morita, Identification of the heparin-binding 111–122.
region of snake venom vascular endothelial growth factor (VEGF-F) and its [156] Y.N. Utkin, A.V. Osipov, Non-lethal polypeptide components in cobra venom, Curr.
blocking of VEGF-A165, Biochemistry 44 (2005) 8858–8864. Pharm. Des. 13 (2007) 2906–2915.
[131] T. Kaji, C. Yamamoto, M. Oh-i, Y. Fujiwara, Y. Yamazaki, T. Morita, A.H. Plaas, T.N. [157] Y. Yamazaki, F. Hyodo, T. Morita, Wide distribution of cysteine-rich secretory pro-
Wight, The vascular endothelial growth factor VEGF165 induces perlecan synthesis teins in snake venoms: isolation and cloning of novel snake venom cysteine-rich
via VEGF receptor-2 in cultured human brain microvascular endothelial cells, secretory proteins, Arch. Biochem. Biophys. 412 (2003) 133–141.
Biochim. Biophys. Acta 1760 (2006) 1465–1474. [158] Y.L. Wang, J.H. Kuo, S.C. Lee, J.S. Liu, Y.C. Hsieh, Y.T. Shih, C.J. Chen, J.J. Chiu, W.G.
[132] T. Nieminen, P.I. Toivanen, N. Rintanen, T. Heikura, S. Jauhiainen, K.J. Airenne, K. Wu, Cobra CRISP functions as an inflammatory modulator via a novel Zn2+-
Alitalo, V. Marjomaki, S. Yla-Herttuala, The impact of the receptor binding profiles and heparan sulfate-dependent transcriptional regulation of endothelial cell adhe-
of the vascular endothelial growth factors on their angiogenic features, Biochim. sion molecules, J. Biol. Chem. 285 (2010) 37872–37883.
Biophys. Acta 1840 (2014) 454–463. [159] S. Lecht, R.A. Chiaverelli, J. Gerstenhaber, J.J. Calvete, P. Lazarovici, N.R. Casewell, R.
[133] M. Simons, E. Gordon, L. Claesson-Welsh, Mechanisms and regulation of endothe- Harrison, P.I. Lelkes, C. Marcinkiewicz, Anti-angiogenic activities of snake venom
lial VEGF receptor signalling, Nat. Rev. Mol. Cell Biol. (2016). CRISP isolated from Echis carinatus sochureki, Biochim. Biophys. Acta 1850 (2015)
[134] F. Duran-Reynals, A spreading factor in certain snake venoms and its relation to 1169–1179.
their mode of action, J. Exp. Med. 69 (1939) 69–81. [160] C.M. Adade, A.L. Carvalho, M.A. Tomaz, T.F. Costa, J.L. Godinho, P.A. Melo, A.P. Lima,
[135] A.F. Wahby, S.M. Mahdy el, H.A. El-Mezayen, W.H. Salama, A.M. Abdel-Aty, A.S. J.C. Rodrigues, R.B. Zingali, T. Souto-Padron, Crovirin, a snake venom cysteine-rich
Fahmy, Egyptian horned viper Cerastes cerastes venom hyaluronidase: purification, secretory protein (CRISP) with promising activity against Trypanosomes and
partial characterization and evidence for its action as a spreading factor, Toxicon Leishmania, PLoS Negl. Trop. Dis. 8 (2014) e3252.
60 (2012) 1380–1389. [161] S.D. Aird, Taxonomic distribution and quantitative analysis of free purine and py-
[136] K.S. Girish, H.P. Mohanakumari, S. Nagaraju, B.S. Vishwanath, K. Kemparaju, Hyal- rimidine nucleosides in snake venoms, Comp. Biochem. Physiol. B Biochem. Mol.
uronidase and protease activities from Indian snake venoms: neutralization by Mi- Biol. 140 (2005) 109–126.
mosa pudica root extract, Fitoterapia 75 (2004) 378–380. [162] A.H. Laustsen, J.M. Gutierrez, B. Lohse, A.R. Rasmussen, J. Fernandez, C. Milbo, B.
[137] K.C.F. Bordon, M.G. Perino, J.R. Giglio, E.C. Arantes, Isolation, enzymatic characteri- Lomonte, Snake venomics of monocled cobra (Naja kaouthia) and investigation
zation and antiedematogenic activity of the first reported rattlesnake hyaluroni- of human IgG response against venom toxins, Toxicon 99 (2015) 23–35.
dase from Crotalus durissus terrificus venom, Biochimie 94 (2012) 2740–2748. [163] A.H. Laustsen, B. Lomonte, B. Lohse, J. Fernández, J.M. Gutiérrez, Unveiling the na-
[138] J. Wohlrab, D. Wohlrab, L. Wohlrab, C. Wohlrab, A. Wohlrab, Use of hyaluronidase ture of black mamba (Dendroaspis polylepis) venom through venomics and anti-
for pharmacokinetic increase in bioavailability of intracutaneously applied sub- venom immunoprofiling: Identification of key toxin targets for antivenom
stances, Skin Pharmacol. Physiol. 27 (2014) 276–282. development, J. Proteome 119 (2015) 126–142.
[139] C.H. Tan, S.Y. Fung, M.K. Yap, P.K. Leong, J.L. Liew, N.H. Tan, Unveiling the elusive [164] C. Delezenne, H. Morel, Action catalytique des venins des serpents sur les acids
and exotic: Venomics of the Malayan blue coral snake (Calliophis bivirgata nucleiques, CR Acad. Sci. 168 (1919) 244–246.
flaviceps), J. Proteome 132 (2016) 1–12. [165] A.R. Taborda, L.C. Taborda, J.N. Williams Jr., C.A. Elvehjem, A study of the
[140] G.A. Wiezel, P.K. dos Santos, F.A. Cordeiro, K.C.F. Bordon, H.S. Selistre-de-Araujo, B. desoxyribonuclease activity of snake venoms, J. Biol. Chem. 195 (1952) 207–213.
Ueberheide, E.C. Arantes, Identification of hyaluronidase and phospholipase B in [166] A.R. Taborda, L.C. Taborda, J.N. Williams Jr., C.A. Elvehjem, A study of the ribonucle-
Lachesis muta rhombeata venom, Toxicon 107 (2015) 359–368. ase activity of snake venoms, J. Biol. Chem. 194 (1952) 227–233.
[141] X. Xu, X.S. Wang, X.T. Xi, J. Liu, J.T. Huang, Z.X. Lu, Purification and partial character- [167] J.G. Georgatsos, M. Laskowski Sr., Purification of an endonuclease from the venom
ization of hyaluronidase from five pace snake (Agkistrodon acutus) venom, Toxicon of Bothrops atrox, Biochemistry 1 (1962) 288–295.
20 (1982) 973–981. [168] S. Iwanaga, T. Suzuki, Enzymes in snake venom, Snake venoms, Springer, 1979
[142] K. Kudo, A.T. Tu, Characterization of hyaluronidase isolated from Agkistrodon 61–158.
contortrix contortrix (Southern Copperhead) venom, Arch. Biochem. Biophys. 386 [169] S.P. Mackessy, Phosphodiesterases, ribonucleases and deoxyribonucleases, En-
(2001) 154–162. zymes from Snake Venoms (1998) 361–404.
[143] Y.H. Mahadeswaraswamy, B. Manjula, S. Devaraja, K.S. Girish, K. Kemparaju, Daboia [170] C.L. Ownby, J. Bjarnason, A.T. Tu, Hemorrhagic toxins from rattlesnake (Crotalus
russelli venom hyaluronidase: purification, characterization and inhibition by beta- atrox) venom. Pathogenesis of hemorrhage induced by three purified toxins, Am.
3-(3-hydroxy-4-oxopyridyl) alpha-amino-propionic acid, Curr. Top. Med. Chem. J. Pathol. 93 (1978) 201–218.
11 (2011) 2556–2565. [171] A.W. Bernheimer, B. Rudy, Interactions between membranes and cytolytic pep-
[144] T.C. Antunes, K.M. Yamashita, K.C. Barbaro, M. Saiki, M.L. Santoro, Comparative tides, Biochim. Biophys. Acta 864 (1986) 123–141.
analysis of newborn and adult Bothrops jararaca snake venoms, Toxicon 56 [172] C.E. Nunez, Y. Angulo, B. Lomonte, Identification of the myotoxic site of the Lys49
(2010) 1443–1458. phospholipase A2 from Agkistrodon piscivorus piscivorus snake venom: synthetic
J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838 837

C-terminal peptides from Lys49, but not from Asp49 myotoxins, exert membrane- [204] A. Meta, H. Nakatake, T. Imamura, C. Nozaki, K. Sugimura, High-yield production
damaging activities, Toxicon 39 (2001) 1587–1594. and characterization of biologically active recombinant aprotinin expressed in Sac-
[173] D. Ma, A. Armugam, K. Jeyaseelan, Cytotoxic potency of cardiotoxin from Naja charomyces cerevisiae, Protein Expr. Purif. 66 (2009) 22–27.
sputatrix: development of a new cytolytic assay, Biochem. J. 366 (2002) 35–43. [205] A. Ritonja, B. Meloun, F. Gubensek, The primary structure of Vipera ammodytes
[174] B.L. Dhananjaya, C.J.M. D'Souza, The pharmacological role of nucleotidases in snake venom trypsin inhibitor I, Biochim. Biophys. Acta 748 (1983) 429–435.
venoms, Cell Biochem. Funct. 28 (2010) 171–177. [206] B. Gocmen, P. Heiss, D. Petras, A. Nalbantsoy, R.D. Sussmuth, Mass spectrometry
[175] K. Trummal, M. Samel, A. Aaspollu, K. Tonismagi, T. Titma, J. Subbi, J. Siigur, E. guided venom profiling and bioactivity screening of the Anatolian Meadow
Siigur, 5'-Nucleotidase from Vipera lebetina venom, Toxicon 93 (2015) 155–163. Viper, Vipera anatolica, Toxicon 107 (2015) 163–174.
[176] M.C. Boffa, G.A. Boffa, Correlations between the enzymatic activities and the factors [207] N.H. Tan, S.Y. Fung, K.Y. Tan, M.K. Yap, C.A. Gnanathasan, C.H. Tan, Functional
active on blood coagulation and platelet aggregation from the venom of Vipera venomics of the Sri Lankan russell's viper (Daboia russelii) and its toxinological
aspis, Biochim. Biophys. Acta 354 (1974) 275–290. correlations, J. Proteome 128 (2015) 403–423.
[177] C. Ouyang, T.F. Huang, Platelet aggregation inhibitors from Agkistrodon acutus [208] D. Petras, P. Heiss, R.D. Sussmuth, J.J. Calvete, Venom proteomics of Indonesian king
snake venom, Toxicon 24 (1986) 1099–1106. cobra, Ophiophagus hannah: Integrating top-down and bottom-up approaches, J.
[178] B.L. Dhananjaya, A. Nataraju, R. Rajesh, C.D. Raghavendra Gowda, B.K. Sharath, B.S. Proteome Res. 14 (2015) 2539–2556.
Vishwanath, C.J. D'Souza, Anticoagulant effect of Naja naja venom 5'-nucleotidase: [209] S.C. Wagstaff, P. Favreau, O. Cheneval, G.D. Laing, M.C. Wilkinson, R.L. Miller, R.
demonstration through the use of novel specific inhibitor, vanillic acid, Toxicon 48 Stöcklin, R.A. Harrison, Molecular characterisation of endogenous snake venom me-
(2006) 411–421. talloproteinase inhibitors, Biochem. Biophys. Res. Commun. 365 (2008) 650–656.
[179] B. Francis, C. Seebart, I.I. Kaiser, Citrate is an endogenous inhibitor of snake venom [210] D. Mora-Obando, J.A. Guerrero-Vargas, R. Prieto-Sanchez, J. Beltran, A. Rucavado,
enzymes by metal-ion chelation, Toxicon 30 (1992) 1239–1246. M. Sasa, J.M. Gutierrez, S. Ayerbe, B. Lomonte, Proteomic and functional profiling
[180] E.D. Rael, Venom phosphatases and 5’-nucleotidases, Enzymes from Snake Venoms of the venom of Bothrops ayerbei from Cauca, Colombia, reveals striking interspe-
1998, pp. 405–423. cific variation with Bothrops asper venom, J. Proteome 96 (2014) 159–172.
[181] Y.H. Chen, T.B. Lo, Chemical studies of formosan cobra (Naja naja atra) venom. Part [211] D. Salazar-Valenzuela, D. Mora-Obando, M.L. Fernandez, A. Loaiza-Lange, H.L.
V. Properties of 5′-nucleotidase, J. Chin. Chem. Soc. 15 (1968) 84–96. Gibbs, B. Lomonte, Proteomic and toxicological profiling of the venom of
[182] J. Dieckhoff, H. Knebel, M. Heidemann, H.G. Mannherz, An improved procedure for Bothrocophias campbelli, a pitviper species from Ecuador and Colombia, Toxicon
purifying 5′-nucleotidase from various sources, Eur. J. Biochem. 151 (1985) 90 (2014) 15–25.
377–383. [212] M. Kohlhoff, M.H. Borges, A. Yarleque, C. Cabezas, M. Richardson, E.F. Sanchez, Ex-
[183] J.M. Gulland, E.M. Jackson, 5-Nucleotidase, Biochem. J. 32 (1938) 597. ploring the proteomes of the venoms of the Peruvian pit vipers Bothrops atrox, B.
[184] N.J. da Silva, S.D. Aird, Prey specificity, comparative lethality and compositional dif- barnetti and B. pictus, J. Proteome 75 (2012) 2181–2195.
ferences of coral snake venoms, Comp. Biochem. Physiol. C Toxicol. Pharmacol. 128 [213] A.K. Tashima, A. Zelanis, E.S. Kitano, D. Ianzer, R.L. Melo, V. Rioli, S.S. Sant'anna, A.C.
(2001) 425–456. Schenberg, A.C. Camargo, S.M. Serrano, Peptidomics of three Bothrops snake
[185] E.A. Zeller, The formation of pyrophosphate from adenosine triphosphate in the venoms: insights into the molecular diversification of proteomes and peptidomes,
presence of a snake venom, Arch. Biochem. 28 (1950) 138–139. Mol. Cell. Proteomics 11 (2012) 1245–1262.
[186] P. Caccin, P. Pellegatti, J. Fernandez, M. Vono, M. Cintra-Francischinelli, B. Lomonte, [214] I. Kregar, P. Locnikar, T. Popovic, A. Suhar, T. Lah, A. Ritonja, F. Gubensek, V. Turk, Bo-
J.M. Gutierrez, F. Di Virgilio, C. Montecucco, Why myotoxin-containing snake vine intracellular cysteine proteinases, Acta Biol. Med. Ger. 40 (1981) 1433–1438.
venoms possess powerful nucleotidases? Biochem. Biophys. Res. Commun. 430 [215] M. Brillard-Bourdet, V. Nguyen, M. Ferrer-di Martino, F. Gauthier, T. Moreau, Puri-
(2013) 1289–1293. fication and characterization of a new cystatin inhibitor from Taiwan cobra (Naja
[187] M. Cintra-Francischinelli, P. Caccin, A. Chiavegato, P. Pizzo, G. Carmignoto, Y. naja atra) venom, Biochem. J. 331 (Pt 1) (1998) 239–244.
Angulo, B. Lomonte, J.M. Gutierrez, C. Montecucco, Bothrops snake myotoxins in- [216] H.J. Evans, A.J. Barrett, A cystatin-like cysteine proteinase inhibitor from venom of
duce a large efflux of ATP and potassium with spreading of cell damage and the African puff adder (Bitis arietans), Biochem. J. 246 (1987) 795–797.
pain, Proc. Natl. Acad. Sci. U. S. A. 107 (2010) 14140–14145. [217] A. Ritonja, H.J. Evans, W. Machleidt, A.J. Barrett, Amino acid sequence of a cystatin
[188] C. Gay, L. Sanz, J.J. Calvete, D. Pla, Snake venomics and antivenomics of Bothrops from venom of the African puff adder (Bitis arietans), Biochem. J. 246 (1987)
diporus, a medically important pitviper in Northeastern Argentina, Toxins (Basel) 799–802.
8 (2015). [218] Y. Hamada, Y. Kiso, New directions for protease inhibitors directed drug discovery,
[189] M. Johnson, M. Kaye, R. Hems, H. Krebs, Enzymic hydrolysis of adenosine phos- Biopolymers 106 (2016) 563–579.
phates by cobra venom, Biochem. J. 54 (1953) 625. [219] X.S. Puente, L.M. Sanchez, C.M. Overall, C. Lopez-Otin, Human and mouse prote-
[190] T.N. Shamsi, R. Parveen, S. Fatima, Characterization, biomedical and agricultural appli- ases: a comparative genomic approach, Nat. Rev. Genet. 4 (2003) 544–558.
cations of protease inhibitors: A review, Int. J. Biol. Macromol. 91 (2016) 1120–1133. [220] M.L. van Hoek, Antimicrobial peptides in reptiles, Pharmaceuticals (Basel) 7 (2014)
[191] L.D. Possani, B.M. Martin, A. Yatani, J. Mochca-Morales, F.Z. Zamudio, G.B. Gurrola, 723–753.
A.M. Brown, Isolation and physiological characterization of taicatoxin, a complex [221] T. Ganz, Defensins: antimicrobial peptides of innate immunity, Nat. Rev. Immunol.
toxin with specific effects on calcium channels, Toxicon 30 (1992) 1343–1364. 3 (2003) 710–720.
[192] N. Willmott, P. Gaffney, P. Masci, A. Whitaker, Novel serine protease inhibitor from [222] R.P. Samy, P. Gopalakrishnakone, S.D. Satyanarayanajois, B.G. Stiles, V.T.K. Chow,
the Australian brown snake, Pseudonaja textilis textilis: inhibition kinetics, Fibrino- Snake venom proteins and peptides as novel antibiotics against microbial infec-
lysis 9 (1995) 1–8. tions, Curr. Proteomics 10 (2013) 10–28.
[193] D. Rodriguez-Ithurralde, R. Silveira, L. Barbeito, F. Dajas, Fasciculin, a powerful an- [223] N.G. de Oliveira Junior, M.H. e Silva Cardoso, O.L. Franco, Snake venoms: attractive
ticholinesterase polypeptide from Dendroaspis angusticeps venom, Neurochem. Int. antimicrobial proteinaceous compounds for therapeutic purposes, Cell. Mol. Life
5 (1983) 267–274. Sci. 70 (2013) 4645–4658.
[194] A.L. Harvey, Twenty years of dendrotoxins, Toxicon 39 (2001) 15–26. [224] B.A. Costa, L. Sanches, A.B. Gomide, F. Bizerra, C. Dal Mas, E.B. Oliveira, K.R. Perez, R.
[195] W.S. Yang, J. Feng, B. Wang, Z.J. Cao, W.X. Li, Y.L. Wu, Z.Y. Chen, BF9, the first func- Itri, N. Oguiura, M.A. Hayashi, Interaction of the rattlesnake toxin crotamine with
tionally characterized snake toxin peptide with kunitz-type protease and potassi- model membranes, J. Phys. Chem. B 118 (2014) 5471–5479.
um channel inhibiting properties, J. Biochem. Mol. Toxicol. 28 (2014) 76–83. [225] T.M. San, J. Vejayan, K. Shanmugan, H. Ibrahim, Screening antimicrobial activity of
[196] J. Fernandez, J.M. Gutierrez, J.J. Calvete, L. Sanz, B. Lomonte, Characterization of a venoms from snakes commonly found in Malaysia, J. Appl. Sci. 10 (2010)
novel snake venom component: Kazal-type inhibitor-like protein from the arbore- 2328–2332.
al pitviper Bothriechis schlegelii, Biochimie 125 (2016) 83–90. [226] A.K. Al-Asmari, R. Abbasmanthiri, N.M. Abdo Osman, Y. Siddiqui, F.A. Al-Bannah,
[197] H. Takahashi, S. Iwanaga, T. Suzuki, Isolation of a novel inhibitor of kallikrein, plas- A.M. Al-Rawi, S.A. Al-Asmari, Assessment of the antimicrobial activity of few
min and trypsin from venom of russell's viper (Vipera russelli), FEBS Lett. 27 (1972) Saudi Arabian snake venoms, Open Microbiol. J. 9 (2015) 18–25.
207-&. [227] M. Hakim, M. Reza, In vitro antibacterial activity of snake venom, Naja naja from
[198] J.J. Calvete, C. Marcinkiewicz, L. Sanz, Snake venomics of Bitis gabonica gabonica. Bangladesh, Br. Biotechnol. J. 8 (2015) 1–5.
Protein family composition, subunit organization of venom toxins, and character- [228] B.L. Ferreira, D.O. Santos, A.L. Dos Santos, C.R. Rodrigues, C.C. de Freitas, L.M. Cabral,
ization of dimeric disintegrins bitisgabonin-1 and bitisgabonin-2, J. Proteome H.C. Castro, Comparative analysis of viperidae venoms antibacterial profile: a short
Res. 6 (2007) 326–336. communication for proteomics, Evid. Based Complement. Alternat. Med. 2011
[199] L. Chang, C. Chung, H.B. Huang, S. Lin, Purification and characterization of a chymo- (2011) 960267.
trypsin inhibitor from the venom of Ophiophagus hannah (king cobra), Biochem. [229] R. Gennaro, M. Zanetti, Structural features and biological activities of the
Biophys. Res. Commun. 283 (2001) 862–867. cathelicidin-derived antimicrobial peptides, Biopolymers 55 (2000) 31–49.
[200] J. Shafqat, Z.H. Zaidi, H. Jornvall, Purification and characterization of a chymotryp- [230] M. Zanetti, Cathelicidins, multifunctional peptides of the innate immunity, J.
sin kunitz inhibitor type of polypeptide from the venom of cobra (Naja naja naja), Leukoc. Biol. 75 (2004) 39–48.
FEBS Lett. 275 (1990) 6–8. [231] V.M. Gomes, A.O. Carvalho, M. Da Cunha, M.N. Keller, C. Bloch Jr., P. Deolindo, E.W.
[201] A.K. Mukherjee, S.P. Mackessy, Pharmacological properties and pathophysiological Alves, Purification and characterization of a novel peptide with antifungal activity
significance of a kunitz-type protease inhibitor (Rusvikunin-II) and its protein from Bothrops jararaca venom, Toxicon 45 (2005) 817–827.
complex (Rusvikunin complex) purified from Daboia russelii russelii venom, [232] M.K. Sachidananda, S.K. Murari, D. Channe Gowda, Characterization of an antibac-
Toxicon 89 (2014) 55–66. terial peptide from Indian cobra (Naja naja) venom, J. Venom. Anim. Toxins Incl.
[202] V. Zupunski, D. Kordis, F. Gubensek, Adaptive evolution in the snake venom kunitz/ Trop. Dis. 13 (2007) 446–461.
BPTI protein family, FEBS Lett. 547 (2003) 131–136. [233] P.G. Correa, N. Oguiura, Phylogenetic analysis of beta-defensin-like genes of
[203] W.M. Chou, W.H. Liu, K.C. Chen, L.S. Chang, Structure-function studies on inhibitory Bothrops, Crotalus and Lachesis snakes, Toxicon 69 (2013) 65–74.
activity of Bungarus multicinctus protease inhibitor-like protein on matrix [234] A.M. Siqueira, N.F. Martins, M.E. De Lima, C.R. Diniz, A. Cartier, D. Brown, B. Maigret,
metalloprotease-2, and invasion and migration of human neuroblastoma SK-N- A proposed 3D structure for crotamine based on homology building, molecular
SH cells, Toxicon 55 (2010) 353–360. simulations and circular dichroism, J. Mol. Graph. Model. 20 (2002) 389–398.
838 J. Boldrini-França et al. / Biochimica et Biophysica Acta 1861 (2017) 824–838

[235] G. Nicastro, L. Franzoni, C. de Chiara, A.C. Mancin, J.R. Giglio, A. Spisni, Solution [268] S. Vaiyapuri, S.C. Wagstaff, K.A. Watson, R.A. Harrison, J.M. Gibbins, E.G.
structure of crotamine, a Na+ channel affecting toxin from Crotalus durissus Hutchinson, Purification and functional characterisation of rhiminopeptidase A, a
terrificus venom, Eur. J. Biochem. 270 (2003) 1969–1979. novel aminopeptidase from the venom of Bitis gabonica rhinoceros, PLoS Negl.
[236] A.M. Torres, P.W. Kuchel, The beta-defensin-fold family of polypeptides, Toxicon Trop. Dis. 4 (2010) e796.
44 (2004) 581–588. [269] S. Zini, M.C. Fournie-Zaluski, E. Chauvel, B.P. Roques, P. Corvol, C. Llorens-Cortes,
[237] L.S. Amer, B.M. Bishop, M.L. van Hoek, Antimicrobial and antibiofilm activity of Identification of metabolic pathways of brain angiotensin II and III using specific
cathelicidins and short, synthetic peptides against Francisella, Biochem. Biophys. aminopeptidase inhibitors: predominant role of angiotensin III in the control of va-
Res. Commun. 396 (2010) 246–251. sopressin release, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 11968–11973.
[238] K. Fosgerau, T. Hoffmann, Peptide therapeutics: current status and future direc- [270] I.M. Francischetti, V. My-Pham, J. Harrison, M.K. Garfield, J.M. Ribeiro, Bitis gabonica
tions, Drug Discov. Today 20 (2015) 122–128. (Gaboon viper) snake venom gland: toward a catalog for the full-length transcripts
[239] I. Schulte, H. Tammen, H. Selle, P. Schulz-Knappe, Peptides in body fluids and tis- (cDNA) and proteins, Gene 337 (2004) 55–69.
sues as markers of disease, Expert. Rev. Mol. Diagn. 5 (2005) 145–157. [271] L. Qinghua, Z. Xiaowei, Y. Wei, L. Chenji, H. Yijun, Q. Pengxin, S. Xingwen, H.
[240] H.M. Doery, J.E. Pearson, Phospholipase B in snake venoms and bee venom, Songnian, Y. Guangmei, A catalog for transcripts in the venom gland of the
Biochem. J. 92 (1964) 599–602. Agkistrodon acutus: identification of the toxins potentially involved in coagulopa-
[241] A.H. Mohamed, A. Kamel, M.H. Ayobe, Studies of phospholipase A and B activities thy, Biochem. Biophys. Res. Commun. 341 (2006) 522–531.
of Egyptian snake venoms and a scorpion toxin, Toxicon 6 (1969) 293–298. [272] H. Ichijo, U. Hellman, C. Wernstedt, L.J. Gonez, L. Claesson-Welsh, C.H. Heldin, K.
[242] A.W. Bernheimer, S.A. Weinstein, R. Linder, Isoelectric analysis of some Australian Miyazono, Molecular cloning and characterization of ficolin, a multimeric protein
elapid snake venoms with special reference to phospholipase B and hemolysis, with fibrinogen- and collagen-like domains, J. Biol. Chem. 268 (1993) 14505–14513.
Toxicon 24 (1986) 841–849. [273] Y. Kakinuma, Y. Endo, M. Takahashi, M. Nakata, M. Matsushita, S. Takenoshita, T.
[243] P. Rey-Suarez, V. Nunez, J. Fernandez, B. Lomonte, Integrative characterization of the Fujita, Molecular cloning and characterization of novel ficolins from Xenopus laevis,
venom of the coral snake Micrurus dumerilii (Elapidae) from Colombia: Proteome, tox- Immunogenetics 55 (2003) 29–37.
icity, and cross-neutralization by antivenom, J. Proteome 136 (2016) 262–273. [274] G. OmPraba, A. Chapeaurouge, R. Doley, K.R. Devi, P. Padmanaban, C. Venkatraman,
[244] S.T. Chatrath, A. Chapeaurouge, Q. Lin, T.K. Lim, N. Dunstan, P. Mirtschin, P.P. D. Velmurugan, Q. Lin, R.M. Kini, Identification of a novel family of snake venom
Kumar, R.M. Kini, Identification of novel proteins from the venom of a cryptic proteins veficolins from Cerberus rynchops using a venom gland transcriptomics
snake Drysdalia coronoides by a combined transcriptomics and proteomics ap- and proteomics approach, J. Proteome Res. 9 (2010) 1882–1893.
proach, J. Proteome Res. 10 (2011) 739–750. [275] A.T. Ching, M.M. Rocha, A.F. Paes Leme, D.C. Pimenta, M. de Fátima D Furtado, S.M.
[245] D.R. Rokyta, A.R. Lemmon, M.J. Margres, K. Aronow, The venom-gland tran- Serrano, P.L. Ho, I.L. Junqueira-de-Azevedo, Some aspects of the venom proteome
scriptome of the eastern diamondback rattlesnake (Crotalus adamanteus), BMC of the Colubridae snake Philodryas olfersii revealed from a Duvernoy's (venom)
Genomics 13 (2012) 312. gland transcriptome, FEBS Lett. 580 (2006) 4417–4422.
[246] M.J. Margres, K. Aronow, J. Loyacano, D.R. Rokyta, The venom-gland transcriptome [276] L. St Pierre, R. Woods, S. Earl, P.P. Masci, M.F. Lavin, Identification and analysis of
of the eastern coral snake (Micrurus fulvius) reveals high venom complexity in the venom gland-specific genes from the coastal taipan (Oxyuranus scutellatus) and re-
intragenomic evolution of venoms, BMC Genomics 14 (2013). lated species, Cell. Mol. Life Sci. 62 (2005) 2679–2693.
[247] S. Flexner, H. Noguchi, Snake venom in relation to haemolysis, bacteriolysis, and [277] M.J. Margres, J.J. McGivern, K.P. Wray, M. Seavy, K. Calvin, D.R. Rokyta, Linking the
toxicity, J. Exp. Med. 6 (1902) 277–301. transcriptome and proteome to characterize the venom of the eastern diamond-
[248] J.W.W. Stephens, W. Myers, The action of cobra poison on the blood: a contribution back rattlesnake (Crotalus adamanteus), J. Proteome 96 (2014) 145–158.
to the study of passive immunity, J. Pathol. Bacteriol. 5 (1898) 279–301. [278] C.R. Diniz, J.M. Gonçalves, Separation of biologically active components from scor-
[249] V. Birdsey, J. Lindorfer, H. Gewurz, Interaction of toxic venoms with the comple- pion venoms by zone electrophoresis, Biochim. Biophys. Acta 41 (1960) 470–477.
ment system, Immunology 21 (1971) 299–310. [279] N.R. Casewell, R.A. Harrison, W. Wüster, S.C. Wagstaff, Comparative venom gland
[250] C.A. Smith, C.W. Vogel, H.J. Muller-Eberhard, Ultrastructure of cobra venom factor- transcriptome surveys of the saw-scaled vipers (Viperidae: Echis) reveal substan-
dependent C3/C5 convertase and its zymogen, factor B of human complement, J. tial intra-family gene diversity and novel venom transcripts, BMC Genomics 10
Biol. Chem. 257 (1982) 9879–9882. (2009) 564.
[251] C.W. Vogel, H.J. Muller-Eberhard, Cobra venom factor: improved method for purifica- [280] S.C. Wagstaff, R.A. Harrison, Venom gland EST analysis of the saw-scaled viper,
tion and biochemical characterization, J. Immunol. Methods 73 (1984) 203–220. Echis ocellatus, reveals novel alpha9beta1 integrin-binding motifs in venom metal-
[252] C.W. Vogel, H.J. Muller-Eberhard, The cobra venom factor-dependent C3 loproteinases and a new group of putative toxins, renin-like aspartic proteases,
convertase of human complement. A kinetic and thermodynamic analysis of a pro- Gene 377 (2006) 21–32.
tease acting on its natural high molecular weight substrate, J. Biol. Chem. 257 [281] A. Chapeaurouge, M.A. Reza, S.P. Mackessy, P.C. Carvalho, R.H. Valente, A. Teixeira-
(1982) 8292–8299. Ferreira, J. Perales, Q. Lin, R.M. Kini, Interrogating the venom of the viperid snake
[253] D.C. Fritzinger, E.C. Petrella, M.B. Connelly, R. Bredehorst, C.W. Vogel, Primary struc- Sistrurus catenatus edwardsii by a combined approach of electrospray and MALDI
ture of cobra complement component C3, J. Immunol. 149 (1992) 3554–3562. mass spectrometry, PLoS One 10 (2015) e0092091.
[254] A.K. Abbas, A.H. Lichtman, S. Pillai, Cellular and molecular immunology, Elsevier, 2012. [282] D. Georgieva, J. Seifert, M. Öhler, M. von Bergen, P. Spencer, R.K. Arni, N. Genov, C.
[255] D.V. Tambourgi, C.W. van den Berg, Animal venoms/toxins and the complement Betzel, Pseudechis australis venomics: adaptation for a defense against microbial path-
system, Mol. Immunol. 61 (2014) 153–162. ogens and recruitment of body transferrin, J. Proteome Res. 10 (2011) 2440–2464.
[256] I.L. Junqueira-de-Azevedo, C.M. Bastos, P.L. Ho, M.S. Luna, N. Yamanouye, N.R. [283] I.L. Junqueira-de-Azevedo, T. Pertinhez, A. Spisni, F.R. Carreno, C.S. Farah, P.L. Ho, Clon-
Casewell, Venom-related transcripts from Bothrops jararaca tissues provide novel ing and expression of calglandulin, a new EF-hand protein from the venom glands of
molecular insights into the production and evolution of snake venom, Mol. Biol. Bothrops insularis snake in E. coli, Biochim. Biophys. Acta 1648 (2003) 90–98.
Evol. 32 (2015) 754–766. [284] J.J. Calvete, P. Juarez, L. Sanz, Snake venomics. Strategy and applications, J. Mass
[257] C.A. Nicolau, P.C. Carvalho, I.L. Junqueira-de-Azevedo, A. Teixeira-Ferreira, M. Spectrom. 42 (2007) 1405–1414.
Junqueira, J. Perales, A.G. Neves-Ferreira, R.H. Valente, An in-depth snake venom [285] B.G. Fry, From genome to "venome": molecular origin and evolution of the snake
proteopeptidome characterization: Benchmarking Bothrops jararaca, J. Proteome venom proteome inferred from phylogenetic analysis of toxin sequences and relat-
(2016). ed body proteins, Genome Res. 15 (2005) 403–420.
[258] C.W. Vogel, D.C. Fritzinger, B.E. Hew, M. Thorne, H. Bammert, Recombinant cobra [286] R. Schoni, The use of snake venom-derived compounds for new functional diag-
venom factor, Mol. Immunol. 41 (2004) 191–199. nostic test kits in the field of haemostasis, Pathophysiol. Haemost. Thromb. 34
[259] C.W. Vogel, D.C. Fritzinger, Humanized cobra venom factor: experimental thera- (2005) 234–240.
peutics for targeted complement activation and complement depletion, Curr. [287] D.C.I. Koh, A. Armugam, K. Jeyaseelan, Snake venom components and their applica-
Pharm. Des. 13 (2007) 2916–2926. tions in biomedicine, Cell. Mol. Life Sci. 63 (2006) 3030–3041.
[260] J. Vejayan, T.L. Khoon, H. Ibrahim, Comparative analysis of the venom proteome of four [288] Y.S. Chan, R.C. Cheung, L. Xia, J.H. Wong, T.B. Ng, W.Y. Chan, Snake venom toxins:
important Malaysian snake species, J. Venom. Anim. Toxins Incl. Trop. Dis. 20 (2014) 6. toxicity and medicinal applications, Appl. Microbiol. Biotechnol. 100 (2016)
[261] S. Li, J. Wang, X. Zhang, Y. Ren, N. Wang, K. Zhao, X. Chen, C. Zhao, X. Li, J. Shao, J. 6165–6181.
Yin, M.B. West, N. Xu, S. Liu, Proteomic characterization of two snake venoms: [289] A.L. Harvey, Toxins and drug discovery, Toxicon 92 (2014) 193–200.
Naja naja atra and Agkistrodon halys, Biochem. J. 384 (2004) 119–127. [290] A.D. Hargreaves, M.T. Swain, M.J. Hegarty, D.W. Logan, J.F. Mulley, Restriction and
[262] G.W. Birrell, S.T. Earl, T.P. Wallis, P.P. Masci, J. de Jersey, J.J. Gorman, M.F. Lavin, The recruitment-gene duplication and the origin and evolution of snake venom toxins,
diversity of bioactive proteins in Australian snake venoms, Mol. Cell. Proteomics 6 Genome Biol. Evol. 6 (2014) 2088–2095.
(2007) 973–986. [291] J. Reyes-Velasco, D.C. Card, A.L. Andrew, K.J. Shaney, R.H. Adams, D.R. Schield, N.R.
[263] Y.F. Pung, S.V. Kumar, N. Rajagopalan, B.G. Fry, P.P. Kumar, R.M. Kini, Ohanin, a Casewell, S.P. Mackessy, T.A. Castoe, Expression of venom gene homologs in di-
novel protein from king cobra venom: its cDNA and genomic organization, Gene verse python tissues suggests a new model for the evolution of snake venom,
371 (2006) 246–256. Mol. Biol. Evol. 32 (2015) 173–183.
[264] S.D. Aird, Snake venom dipeptidyl peptidase IV: taxonomic distribution and quantita- [292] V.J. Lynch, Inventing an arsenal: adaptive evolution and neofunctionalization of
tive variation, Comp. Biochem. Physiol. B Biochem. Mol. Biol. 150 (2008) 222–228. snake venom phospholipase A2 genes, BMC Evol. Biol. 7 (2007) 2.
[265] Y. Ogawa, Y. Mamura, N. Murayama, R. Yanoshita, Characterization and cDNA clon- [293] B.G. Fry, W. Wuster, R.M. Kini, V. Brusic, A. Khan, D. Venkataraman, A.P. Rooney,
ing of dipeptidyl peptidase IV from the venom of Gloydius blomhoffi brevicaudus, Molecular evolution and phylogeny of elapid snake venom three-finger toxins, J.
Comp. Biochem. Physiol. B Biochem. Mol. Biol. 145 (2006) 35–42. Mol. Evol. 57 (2003) 110–129.
[266] Y. Ogawa, M. Kanai-Azuma, Y. Akimoto, H. Kawakami, R. Yanoshita, Exosome-like [294] N.R. Casewell, W. Wuster, F.J. Vonk, R.A. Harrison, B.G. Fry, Complex cocktails: the
vesicles in Gloydius blomhoffii blomhoffii venom, Toxicon 51 (2008) 984–993. evolutionary novelty of venoms, Trends Ecol. Evol. 28 (2013) 219–229.
[267] S.C. Yamasaki, J.S. Villarroel, J.M. Barone, L. Zambotti-Villela, P.F. Silveira, Amino- [295] N.R. Casewell, S.C. Wagstaff, R.A. Harrison, C. Renjifo, W. Wuster, Domain loss facil-
peptidase activities, oxidative stress and renal function in Crotalus durissus itates accelerated evolution and neofunctionalization of duplicate snake venom
terrificus envenomation in mice, Toxicon 52 (2008) 445–454. metalloproteinase toxin genes, Mol. Biol. Evol. 28 (2011) 2637–2649.

S-ar putea să vă placă și